You are on page 1of 23

Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

Contents lists available at ScienceDirect

Neuroscience and Biobehavioral Reviews


journal homepage: www.elsevier.com/locate/neubiorev

The effect of high fat, high sugar, and combined high fat-high sugar diets on T
spatial learning and memory in rodents: A meta-analysis
Kirsten N. Abbotta, Christopher K. Arnottb, R. Frederick Westbrooka, Dominic M.D. Tranb,

a
School of Psychology, The University of New South Wales, Australia
b
School of Psychology, The University of Sydney, Australia

ARTICLE INFO ABSTRACT

Keywords: Evidence from human and animal studies suggests that high-energy diets impair cognitive function. However,
Rodent models the conditions for such impairments are unclear as studies have differed in the type and duration of diet exposure
High energy diet as well as in the tasks used to assess deficits in cognition. Here, we focused on hippocampal-dependent tasks. We
High fat diet conducted separate meta-analyses of the results from rodent studies using: 1) different diets (high in fat, high in
High sugar diet
sugar, or high in both fat and sugar); and 2) different tasks to assess hippocampal-dependent spatial learning and
Hippocampus
Spatial learning and memory
memory (water maze, place recognition, radial arm maze, and spontaneous alternation). We focused on the
effects of relatively short-term dietary manipulations and, therefore, restricted our analyses to studies that
provided the diet for 2 months or less. The meta-analyses showed that each type of diet and task adversely
affected performance, with the largest effect produced by exposure to a combined high fat-high sugar diet and
the use of the radial arm maze to assess the effect of such diets on cognition.

1. Introduction called “fast” foods predicted scores on subtests of the Wechsler Memory
Scale Revised (WMS-R; Wechsler and Stone, 1987) that are sensitive to
The modern diet contains many foods and drinks that are rich in left hippocampal damage, specifically, the verbal paired associates and
saturated fat and sugar (Kearney, 2010; Khatibzadeh et al., 2016; logical memory subtests. However, intake of such foods failed to predict
Ritchie and Roser, 2017; Vasileska and Rechkoska, 2012; reviewed in scores on the hippocampal-independent, digit span subtest of the WMS-
Cordain et al., 2005; Popkin et al., 2012). It has long been recognized R. Similarly, cross-sectional research in healthy young women found
that excessive intake of this diet is associated with increased body that habitual intake of saturated fat predicted lower scores on the
weight, even obesity, some forms of cancer, as well as metabolic and CANTAB eclipse (v3) suite of cognitive tests (Cambridge Cognition Ltd.,
cardiovascular diseases (Carrera-Bastos et al., 2011; Cordain et al., 2006) that are sensitive to hippocampal damage, such as verbal re-
2005; Mente et al., 2009). More recent epidemiological evidence has cognition memory, delayed-matching-to sample, and paired associate
indicated that this diet is also associated with deficits in cognition learning (Gibson et al., 2013). Finally, there is even evidence that
(reviewed in Kanoski and Davidson, 2011, and Yeomans, 2017). For cognitive deficits can be detected after relative short exposures to such
example, longitudinal studies of middle-aged (Eskelinen et al., 2008) diets. Attuquayefio et al. (2017) found a statistically significant pre- to
and older adults (Gustaw-Rothenberg, 2009; Morris et al., 2004; post decline in scores on the Hopkins Verbal Learning Task Revised in
Okereke et al., 2012; Naqvi et al., 2011; Roberts et al., 2012) showed healthy adults who consumed a breakfast high in saturated fat and
that consumption of diets high in saturated fat and refined carbohy- refined sugar for four days, but not in a control group of adults that
drates was negatively correlated with overall cognitive “well-being” consumed a low fat, low sugar diet of similar palatability.
(e.g., memory, executive function, psychomotor speed) in later life and Rodent models have confirmed that diets high in saturated fat,
positively correlated with neurological disorders such as Alzheimer’s sugar, or both fat and sugar are associated with deficits in tasks that
disease (reviewed in Pugazhenthi et al., 2017). require the hippocampus and surrounding cortices (see Table 1). Sev-
Hippocampal-dependent forms of cognition seem to be particularly eral studies have reported such dietary effects in the use of spatial cues
susceptible to the detrimental effects of diets that are high in saturated to navigate to the hidden platform in the Morris (e.g., Che et al., 2018)
fat and/or refined sugar. A cross-sectional study of young adults and radial arm water mazes (e.g., Alzoubi et al., 2013a, b; Alzoubi
(Francis and Stevenson, 2011) found that self-reported intake of so- et al., 2018), or to avoid arms that have been visited and depleted of


Corresponding author.
E-mail address: minh.d.tran@sydney.edu.au (D.M.D. Tran).

https://doi.org/10.1016/j.neubiorev.2019.08.010
Received 11 February 2019; Received in revised form 2 August 2019; Accepted 12 August 2019
Available online 24 August 2019
0149-7634/ © 2019 Elsevier Ltd. All rights reserved.
Table 1
Body weight and metabolic parameters of experimental and control animals provided a high fat, high sugar, or high fat-high sugar diet (* = unless otherwise stated; HFD = high fat diet; HSD = high sugar diet;
HFHS = high fat high sugar diet; CD = control diet; MWM = Morris Water Maze; PR = Place recognition; RAM = Radial arm maze; RAWM = Radial Arm Water Maze; SA = Spontaneous Alternation; RI = Retention
Interval; GTT = Glucose Tolerance Test; ITT = Insulin Tolerance Test; TC = Total Cholesterol; HDL = High Density Lipoprotein; LDL = low Density Lipoprotein; ELISA = Enzyme Linked Immunosorbent Assay; HOMA
K.N. Abbott, et al.

= Homeostatic Model Assessment; QUICKI = Quantitative Insulin Sensitivity Check).


Study Species & Experimental Diet Spatial Task Diet-induced Body Weight (BW) Blood* Glucose Level Plasma* Insulin Level HOMA Index Plasma* Others
Strain [Sex & Classification for Meta- Deficit (BGL) Triglycerides
Age] Analysis

Abbott et al (2016) Expt Sprague HSD PR Yes No effect – – – – –


1 Dawley rats Object-in- No (Female);
[Male & Place Yes (Male)
Female]
Alzoubi et al (2013a) Wistar rats HFD (labelled Western Diet RAWM (6- Yes Sig effect – – – – –
Expt 1 [Male] by authors) arm) (P < .05)
Alzoubi et al (2013b) Wistar rats HFD (labelled High Fat RAWM (6- Yes Sig effect – – – – –
Expt 1 [Male] Cholesterol Diet by authors) arm) (P < .05)
Alzoubi et al (2018) Wistar rats HFD RAWM (6- Yes – – – – – –
[Male] arm)
Arnold et al (2014) C57BL/6 J Extreme HFD T-Maze SA Yes Sig effect Sig effect (P < .001) – – – –
mice [Male] (P < .01)
Moderate HFD T-Maze SA Yes Sig effect Sig effect on BGL and – – – –
(P < .05) GTT response
(Ps < .05)
Ayabe et al (2018) C57BL/6 J HFD PR No Sig effect [week – – – Sig effect –
mice [Male] 3+] (Ps < .01) (P < .01)
Beilharz et al (2014) Sprague HFHS (labelled Cafeteria PR Yes (all tests) Sig effect [day – – – – –
Expt 1 Dawley rats Diet with Sucrose solution by 16+] (P < .05)
[Male] authors)

400
Beilharz et al (2014) Sprague HSD PR Yes (all tests) No effect – [ELISA] no effect – No effect [ELISA] no effect on
Expt 2 Dawley rats plasma leptin
[Male] HFHS (labelled Cafeteria Yes (all tests) Sig effect [day – [ELISA] sig effect Sig effect [ELISA] sig effect
Diet with Sucrose solution by 15+] (Ps < .05); (P < .05) (P < .05) plasma leptin
authors) (P < .05)
HFHS (labelled Cafeteria Yes (all tests) Sig effect [day – [ELISA] no effect Sig effect [ELISA] sig effect
Diet by authors) 15+] (Ps < .05); (P < .05) plasma leptin
(P < .05)
Beilharz et al (2018) Sprague HFHS (labelled Cafeteria PR Yes (all tests) Sig effect No effect – – – –
Dawley rats Diet with Sucrose solution by (P < .001)
[Male] authors)
Boitard et al (2014) Wistar rats HFD MWM No Sig effect Sig effect (P < .05) No effect – No effect No effect on plasma
[Male] (P < .05) cholesterol; sig effect
plasma leptin
(P < .05)
Che et al (2018) SAMP8 mice HFD MWM Yes – – – – – –
[Male]
Denver et al (2018) C57BL/6 J HFD 10 days MWM No Sig effect [day 4+] – – – – –
mice [Male] (Ps < .05)
HFD 26 days No Sig effect [day – – – – –
24+] (P < .05)
HFD 8 weeks No Sig effect [week – – – – –
2+] (Ps < .05)
Gergerlioglu et al (2016) Wistar rats HFD MWM Yes No effect No effect – – – –
Expt 1 [Male] HSD No No effect No effect – – – –
Hargrave et al. (2016) Sprague Ketogenic Diet Y-maze SA No No effect No effect No effect – – –
Expt 1 Dawley rats HFHS(labelled high-fat, Y-maze SA Yes (10 day No effect No effect No effect – – –
[Male] high-dextrose “Western” diet test); No (40
by authors) day test)
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

(continued on next page)


Table 1 (continued)

Study Species & Experimental Diet Spatial Task Diet-induced Body Weight (BW) Blood* Glucose Level Plasma* Insulin Level HOMA Index Plasma* Others
Strain [Sex & Classification for Meta- Deficit (BGL) Triglycerides
Age] Analysis
K.N. Abbott, et al.

Jamshed et al (2014) Sprague HFD (labelled cholesterol- MWM Yes – – – – Sig effect week 6 Sig effect on TC, LDL,
Expt 1 Dawley rats chocolate-butter fat diet by (P < .01) & HDL:LDL ratio in
[Male & authors) week 6 (Ps < .01)
Female]
Jurdak et al (2008) Long– Evans HFD MWM No Sig effect Overnight Fasted. No – – No effect No effect on TC, HDL,
rats [Male] (P < .05) effect BGL or GTT or non-HDL.
response
HSD No (24-hr RI); Sig effect Overnight Fasted. Sig – – Sig effect No effect on TC, HDL,
Yes (10-day RI) (P < .05) effect BGL (P < . 01) (P < .05) or non-HDL.
but not GTT response
Kanoski and Davidson Sprague HFHS (labelled High Energy RAM Yes Sig effect [day – – – – –
(2010) Dawley rats Diet by authors) (8 ar m) 14+] (Ps < .05)
[Male]
Kendig et al (2013) Hooded Wistar HSD MWM Yes (Probe No effect 6 -hour Fasted. Sig – – – –
rats [Male & tests 1-3); No effect [day 29+]
Female] (28-day RI) (main effect,
P < .05) but not day
14
Liang et al (2015) Sprague HFD MWM Yes – – – – – –
Dawley rats
[Male]
Magnusson et al (2015) C57BL/6 J HFD MWM No Sig effect [week – – – – –
mice [Male] MWM Yes 4+] (Ps < .05).
Reversal

401
Task
PR No
HSD MWM No No effect – – – – –
MWM Yes
Reversal
Task
PR No
Mirzaei et al (2018) Wistar rats HFD MWM Yes – – – – – –
[Male]
Molteni et al (2004) Fisher 344 rats Powdered HFHS MWM Yes – – – – – –
[Female]
Moreira et al (2014) Swiss Albino HFD (labelled PR Yes No effect No effect on plasma – Sig effect Sig effect on TC, non-
Mice [Male] Hypercholestermic Diet by glucose (P < .05) HDL, & HDL
authors) (Ps < .05)
Park et al (2018) ICR mice HFD Y-Maze SA Yes Sig effect [week – – – – –
[Male] 5+] (Ps < .05)
Pathan et al (2008) Sprague HFD MWM Yes Sig effect [ELISA] sig effect [ELISA] sig effect – [ELISA] sig effect [ELISA] sig effect on
Dawley rats (P < .05) (P < .001) (P < .001) on blood trigs blood TC (P < .001)
[Male] (P < .001)
Pratchayasakul et al Wistar rats HFD MWM No No effect 12 -hour Fasted. No 12 -hour Fasted. No effect 12 -hour Fasted. 12 -hour Fasted. No
(2015) [Female] 4 effect on plasma [ELISA] no effect No effect effect on TC, LDL, or
weeks old glucose or GTT HDL.
response
Wistar rats No Sig effect 12 -hour Fasted. No 12 -hour Fasted. Sig effect 12 -hour Fasted. 12 -hour Fasted. Sig
[Female] 8 (P < .05) effect on plasma [ELISA] sig effect (P < .05) No effect effect on TC, & LDL
weeks old glucose or GTT (P < .05) (Ps < .01) but not
response HDL
(continued on next page)
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
Table 1 (continued)

Study Species & Experimental Diet Spatial Task Diet-induced Body Weight (BW) Blood* Glucose Level Plasma* Insulin Level HOMA Index Plasma* Others
Strain [Sex & Classification for Meta- Deficit (BGL) Triglycerides
Age] Analysis
K.N. Abbott, et al.

Ross et al (2013) Sprague HFHS PR No Sig effect – – – – –


Dawley rats (P < .05)
[Male]
Scichilone et al (2016) Sprague Ketogenic Diet MWM No No effect Sig effect week 2 – – – –
Dawley rats (P < .01), but not
[Male] week 4 or 8
Silva et al (2005) Wistar rats Ketogenic Diet MWM No – – – – – –
[Male]
Spencer et al (2017) F344xbn F1 HFD MWM No Sig effect No effect [ELISA] no effect – – –
rats [Male] 3- (P < .05)
months old
F344xbn F1 No Sig effect No effect [ELISA] no effect – – –
rats [Male] 24- (P < .05)
months old
Spinelli et al (2017) C57BL/6 J HFD MWM Yes – – [ELISA] sig effect – – –
mice [Male] (P < .05)
Takechi et al (2017) C57BL6/J mice High Fat Fructose Diet MWM No Sig effect Sig effect on blood [ELISA] sig effect Non-fasted. – –
[Male] (P < .001) glucose and hba1c (P < .05) Sig effect
(P < .01) (P < .01)
Thirumangalakudi et al C57BL/6 mice HFD RAWM Yes – – – – – –
(2008) (8 ar m)
Tran and Westbrook Sprague HFHS PR Yes (all tests) Sig effect overall – – – – –
(2015) Expt 1 Dawley rats (P < .001) but not
[Male] at first test (day 6-

402
7)
Tran and Westbrook Sprague HFHS PR Yes – – – – – –
(2018) Expt 1 Dawley rats
[Male]
Valladolid-Acebes et al C57BL/6 J HFD RAM Yes Sig effect [week 6 -hour Fasted. No 6 -hour Fasted sig – 6-hour Fasted. No 6 -hour Fasted. No
(2011) mice [Male] (8 ar m) 2+] (Ps < .001) effect effect (P < .001) effect effect on plasma
leptin
Valladolid-Acebes et al C57BL/6 J HFD PR Yes Sig effect No effect No effect – – Sig effect plasma
(2013) Expt 1 mice [Male] (P < .001) leptin (P < .001)
Valladolid-Acebes et al PR No Sig effect Sig effect (P < .05) No effect – – Sig effect plasma
(2013) Expt 3 (P < .001) leptin (P < .001)
Wei et al (2018) C57BL/6 J HFD MWM Yes Sig effect [week – – – – –
mice [Male] 2+] Ps < .05)
Woodie and Blythe Sprague HFD MWM No Sig effect (BW and 4 - 6 hour Fasted. No 4 - 6 hour Fasted. – 4 - 6 hour Fasted. 4-6 hour Fasted. No
(2018) Dawley rats MWM No % BW change; effect [ELISA] no effect vs No effect effect on QUICKI
[Male] Reversal Ps < .05); CD; sig effect vs
Task HfruD (P < .05).
High Fructose Diet MWM No No effect (BW and 4 - 6 hour Fasted. No 4 - 6 hour Fasted. – 4 - 6 hour Fasted. 4-6 hour Fasted. No
MWM No % BW change) effect. [ELISA] no effect vs No effect effect on QUICKI
Reversal CD
Task
Wu et al (2003) Expt 1 Sprague Powdered HFHS MWM No – – – – – –
Dawley rats
[Male]
Xu and Reichelt (2018) Sprague HSD PR Yes No effect – – – – –
Dawley rats
[Male]
Xu et al (2018) C57BL/6 J HFD MWM Yes Sig effect [week Sig effect on BGL and 4 -hour Fasted. Sig – Sig effect on serum Sig effect of serum
mice [Male] 2+] (Ps < .05) GTT response effect on ITT response trigs (P < .05) TC, LDL, & HDL
(Ps < .01) (Ps < .01) (Ps < .01)
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

food in the radial arm maze (e.g. Kanoski and Davidson, 2010). Other criteria for inclusion in analysis. First, we only included studies in
studies have found deficits in short-term place, but not object re- which rodents were provided with the diet for a maximum of eight
cognition memory (e.g., Abbott et al., 2016; Beilharz et al., 2014; Tran weeks (or two months). We did so on the assumption that such rela-
and Westbrook, 2015), and in spontaneous alternation tasks in Y- tively short dietary manipulations would be more likely to allow us to
(Hargrave et al., 2016; Park et al., 2018) and T-mazes (Arnold et al., detect any differential effects of the diet composition and/or behavioral
2014). Deficits in these tasks have been observed in rodents fed such task used for assessment of hippocampal-dependent learning and
diets across time periods that have ranged from months (e.g., Molteni memory. Inspection of the literature indicated that eight weeks of
et al., 2002) to weeks (Abbott et al., 2016; Hargrave et al., 2016), and dietary exposure was both short enough to reveal differences and long
even days (Beilharz et al., 2014; Kanoski and Davidson, 2010; Murray enough to provide a sufficient sample for analysis. We note that some
et al., 2009; Tran and Westbrook, 2015), indicating that deficits can studies included in the meta-analysis report disturbances in metabolic
occur rapidly, well in advance of diet-induced increases in body weight. parameters, such as hyperglycemia, insulin resistance, and body weight
Although diet-induced deficits in cognition are relatively well- gain (see Table 1 and Discussion), but did not incorporate these dis-
documented, there have also been failures to detect such effects. For turbances into the analyses due to the lack of statistical power arising
example, rats fed a high fat-high sugar diet for four weeks (Ross et al., from the large variability in the parameters measured and reported
2013) or mice fed a high fat diet for seven weeks (Ayabe et al., 2018) across studies. The second inclusion criterion was studies that used
did not differ significantly from chow-fed controls in the place re- rodents older than four weeks. We did so in order to ensure that the
cognition memory task at four hour (or longer) retention intervals. animals were weaned, which typically occurs between postnatal day 21
Further, several investigators (Jurdak et al., 2008; Pratchayasakul et al., to 28 (Sengupta, 2011, 2013), and to avoid any interaction effects of
2015; Scichilone et al., 2016; Silva et al., 2005; Zhao et al., 2004) failed dietary interventions on developmental processes, such as hippocampal
to detect differences in spatial memory in the Morris water maze be- maturation, which occurs at approximately postnatal day 21 (Fan et al.,
tween rats fed a high fat diet for four to eight weeks and chow-fed 2018; Radic et al., 2017; Semple et al., 2013).
control. Finally, older (eight week old) mice placed on a high fat diet for
eight weeks failed to show a deficit in place recognition memory at a 2. Method
one hour retention interval, whereas younger (five week old) placed on
that diet for that length of time did show a deficit in place recognition 2.1. Eligibility criteria
memory (Valladolid-Acebes et al., 2013).
A major difficulty in interpreting the discrepant findings is due to The meta-analysis was limited to published, peer-reviewed research
the procedural variation across studies, especially in the dietary ma- that fed rats or mice diets high in either fat or sugar, or high in both fat
nipulations and the behavioral tasks used to assess the effects of these and sugar. To be included in the analysis, the animals were required to
manipulations. One of the most notable differences between studies is be at least 4 weeks of age at the start of diet administration and be on
diet composition. For example, lard is typically used to increase the the diet for no longer than 8 weeks (or 2 months). For studies that used
amount of fat in the diet (Che et al., 2018; Kang et al., 2014; Wei et al., transgenic rodent models, the only data included for analysis were from
2018), but saturated fatty acids (Moreira et al., 2014; Murray et al., control and diet intervention wild type animals. For studies that in-
2009; Spinelli et al., 2017), hydrogenated vegetable oil (Jurdak et al., cluded a drug manipulation, the only data included for analysis was
2008), and suet (Gergerlioglu et al., 2016) have also been used. The taken from control and diet intervention animals that received saline
total content of fat has also varied between studies, ranging from 18 g/ administration. Studies that used a maternal diet manipulation were
kg (Che et al., 2018) to approximately 47.5 g/kg of lard (Scichilone excluded. All the included studies used a hippocampal-dependent task,
et al., 2016; Zhao et al., 2004). Sugar content also varies across studies, specifically, the Morris water maze (Compton et al., 1997), radial arm
ranging from sucrose solution concentrations of 10% (Abbott et al., maze and radial arm water maze (Floresco et al., 1997), place and
2016; Kendig et al., 2013) to 32% (Jurdak et al., 2008). Finally, other object-in-place recognition memory (Barker and Warburton, 2011), and
studies have altered the sugar content of the solid foods in the diet spontaneous alternation (Lalonde, 2002; Stevens and Cowey, 1973).
(Gergerlioglu et al., 2016; Magnusson et al., 2015). A second notable Some studies have used other hippocampal-dependent tasks (e.g.,
difference is the task used to assess cognitive ability. These tasks in- context fear conditioning; Sobesky et al., 2014), but these studies did
clude the various types of mazes and recognition memory paradigms. not meet the other inclusion criteria. A number of studies have in-
Moreover, there are between-study differences in the parameters used vestigated the effects of diet on occasion-setting/conditioned inhibition
in each of these tasks. For example, studies using the Morris water maze processes (e.g., hippocampal-dependent feature negative discrimina-
differed in the number of acquisition trials, the duration of the interval tion) but these either did not meet the inclusion criterion of diet ex-
between trials as well as that between the final acquisition trial and the posure duration (Kanoski et al., 2007, 2010), or were excluded for
first test trial (where the hidden platform is absent or moved to a new splitting animals into diet-resistant and diet-prone obesity groups
location). (Davidson et al., 2013).
The several failures to replicate reports of diet-induced deficits in
hippocampal-dependent forms of learning and memory raise the pos- 2.2. Database search
sibility that such deficits are relative fragile. Alternatively, such deficits
may be more readily induced by certain dietary manipulations and/or We used Pubmed and Proquest central databases (Appendix A) and
the tasks and protocols used to assess their effects. While this topic has limited the search to papers published on or before June 30th 2018.
been extensively reviewed (see Beilharz et al., 2015; Cordner and
Tamashiro, 2015; Kanoski and Davidson, 2011; Yeomans, 2017), as far 2.3. Screening for inclusion
as we are aware, an analysis of the circumstances which do or do not
produce diet-induced deficits in cognition has not been conducted. We The process of selecting papers for inclusion in the meta-analysis is
aimed to provide such an analysis by subjecting some of the published shown in Fig. 1. Papers were first screened for eligibility using in-
literature on diet-induced deficits in cognition to a meta-analysis. More formation provided in the title and abstract. Full texts of papers that
specifically, we sought, first, to assess the contribution of diets high in passed the eligibility test were then screened by two reviewers to
saturated fat, refined sugar, or both, to impairments of spatial learning confirm that they had in fact met the requirements for inclusion in the
and memory, and, second, to examine whether inconsistent findings in meta-analysis. In instances where there was disagreement, a third party
the literature were due to differences in the composition of the diet would have been asked to decide whether an article reached the elig-
and/or the behavioral task used to assess its effects. We used two ibility criteria, but there was no such disagreement.

403
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

analysis because the division into subgroups by three diet conditions


and four task types had a high degree of variance in the number of
experiments (or datasets) included in each subgroup. Additionally, the
datasets in one subgroup were not all strictly independent from those in
another subgroup; some datasets included across subgroups used the
same control group for comparison (e.g., for studies that compared
multiple diet manipulations to a single control group, the values for the
single control group were used in each of the diet manipulation meta-
analyses). Regardless, we were not so much concerned with an omnibus
test result from a subgroups meta-analysis but rather with the magni-
tude of the effect size in each subgroup. Therefore, we ran separate
meta-analyses with a Bonferroni correction to control for the use of the
same data set across two analyses (by diet and by task), an alpha of
α = 0.025 was considered significant.

3. Results

3.1. Study selection

The initial search yielded 991 articles. After the removal of dupli-
cates, 797 articles were screened based on the title and abstract, and
717 were excluded. Of the remaining 80 articles reviewed in full text, a
further 39 were excluded as they did not meet inclusion criteria. Of
these articles, 14 had a diet manipulation longer than the eight weeks
prior to behavioral testing, nine did not include a spatial learning and
memory behavioral task, seven started diet manipulations in animals
younger than 4 weeks, two had transgenic subjects but no wild type
control, two involved drug manipulations but lacked a vehicle control,
Fig. 1. Flow chart showing the process of selecting articles for inclusion in the
meta-analyses. two did not make statistical comparisons between groups, one had no
appropriate control diet condition, one assessed the effect of maternal
diet on offspring, and one was a review article.
2.4. Data collection
3.2. Study characteristics
The two reviewers independently extracted data from studies that
met the eligibility criteria and recorded information about the animals, Across the 41 articles that met eligibility criteria (with some articles
diet manipulation, behavioral assays, outcomes measured, and results. comprising of multiple eligible experiments or datasets), there were a
For studies where subject body weights were provided, growth charts total of 32 experiments that investigated the effect of a high fat diet
from Charles River Laboratories were used to confirm animals were at (Table 2), nine experiments investigated the effect of a high sugar diet
least 4 weeks of age based on weights provided. For studies where there (Table 3), and 11 experiments investigated the effect of a diet high in
was no statistically significant difference between control and diet an- both fat and sugar (Table 4). The effect of diet on spatial learning and
imals but insufficient statistical details were provided in the published memory was assessed via the use of: extra-maze cues to navigate in the
methods and supplementary material, the corresponding author was Morris water maze, radial arm maze, or radial arm water maze; spon-
contacted twice to provide these details. If details were not provided, a taneous alternation in a T- or Y-maze; or via the use of novel object
nominal F- or t-value of 1 was used for the meta-analysis. Excluding location in recognition memory tasks. Twenty-two experiments used
these results would have contributed to any publication biases towards the Morris water maze (Table 5), 13 used the place recognition task
papers reporting statistically significant results. When the number of (Table 6), six used the radial arm or radial arm water maze (Table 7),
animals used for statistical analysis was provided as a range rather a and four used spontaneous alternation (Table 8).
single value, the lowest, most conservative numerical value of the range
was used for analysis of effect size. When a study had multiple test 3.2.1. High fat diet
points but did not report a main effect, the data from each individual A total of 674–706 rats and mice (control n = 335–351; high fat diet
test point was included. For studies that compared multiple diet ma- n = 339–355) were used to assess the effect of a high fat diet intake. Of
nipulations to a single control group, the values for the single control the 32 studies, 13 used C57BL/6 J mice (Arnold et al., 2014; Ayabe
group were used in each of the diet manipulation meta-analyses. et al., 2018; Denver et al., 2018; Magnusson et al., 2015; Silva et al.,
2005; Spinelli et al., 2017; Thirumangalakudi et al., 2008; Valladolid-
2.5. Data analysis Acebes et al., 2011, 2013; Wei et al., 2018; Xu et al., 2018), eight used
Wistar rats (Alzoubi et al., 2013a, 2013b, 2018; Boitard et al., 2014;
Data were analyzed using Comprehensive Meta-Analysis version 3 Gergerlioglu et al., 2016; Mirzaei et al., 2018; Pratchayasakul et al.,
(CMA v3). The studies were organized by (1) type of diet (high fat, high 2015; Silva et al., 2005), and six used Sprague-Dawley rats (Hargrave
sugar, and high fat-high sugar) and (2) type of behavioral task (Morris et al., 2016; Jamshed et al., 2014; Liang et al., 2015; Pathan et al.,
water maze, place recognition, radial arm maze, and spontaneous al- 2008; Scichilone et al., 2016; Woodie and Blythe, 2018). The remaining
ternation). The standardized mean difference (SMD), calculated as studies used Swiss Albino mice (Moreira et al., 2014), ICR mice (Park
Hedge’s g (Hedges, 1981) and weighted by sample size, was used to et al., 2018), SAMP-8 mice (Che et al., 2018), F344xBN F1 rats (Spencer
measure the effects of diet and task. A random effects model was se- et al., 2017), and Long-Evans rats (Jurdak et al., 2008). All the animals
lected due to differences across the studies in calories, percentage of were males, except for Jamshed et al. (2014) who used males and fe-
fats and sugars, and behavioral tests. Therefore, statistics measuring males, and Pratchayasakul et al. (2015) who used only females.
heterogeneity are not reported. We did not run a subgroups meta- A total of 53 dependent variables were extracted from the 32

404
Table 2
Composition and energy density of experimental and control diets provided to animals to examine the effect of high fat diet (HFD) consumption on hippocampal-dependent learning and memory. (* = where provided;
w/w = weight / weight; kcal = kilocalories).
Author & Expt Species & Strain [sex] Experimental diet [Energy Density*] Experimental diet Label from Fat Content Protein Control Diet [Energy Density*]
K.N. Abbott, et al.

Authors

Alzoubi et al (2013a) Expt 1 Wistar rats [Male] Formulated diet (Sahil-Huran Animal Food Western Diet 25% (w/w) total fat including 18% (w/w) Standard diet (Sahil-Huran Animal
Company, Jordan) 11% (w/w) Unsat fat Food Company)
Alzoubi et al (2013b) Expt 1 Wistar rats [Male] Formulated diet (Sahil-Huran Animal Food High Fat Cholesterol Diet 25% (w/w) total fat including 18% (w/w) Standard diet (Sahil-Huran Animal
Company, Jordan) 11% (w/w) Unsat fat Food Company)
Alzoubi et al (2018) Wistar rats [Male] Formulated diet (Unknown supplier) HFD 25% (w/w) total fat including 18% (w/w) Standard diet (Sahil-Huran Animal
11% (w/w) Unsat fat Food Company)
Arnold et al (2014) C57BL/6 J mice [Male] Research Diets #D12492 [5.21 kcal/g] Extreme HFD 60% kcal 20% kcal Research Diets Low-Fat Rodent Diet
#D12450B [3.82 kcal/g]
Research Diets #D12451 [4.7 kcal/g] Moderate HFD 45% kcal 20% kcal Research Diets Low-Fat Rodent Diet
#D12450B [3.82 kcal/g]
Ayabe et al (2018) C57BL/6 J mice [Male] Research Diets #D12492 [5.21 kcal/g] HFD 60% kcal (30% w/w Lard) 20% kcal Research Diets Low-Fat Rodent Diet
#D12450 J [3.85 kcal/g]
Boitard et al (2014) Wistar rats [Male] Research Diets #D12451 [4.7 kcal/g] HFD 24% (w/w) – Standard chow A04 SAFE (Augy,
France) [2.9 kcal/g]
Che et al (2018) SAMP8 mice [Male] OpenSource Diets #D12492 HFD 18.4% (w/w) Lard + 4.6% 14% OpenSource Diets Low Fat Chow
(w/w) Corn Oil #D12450
Denver et al (2018) C57BL/6 J mice [Male] Special Diet Services (Witham, UK) HFD 45% (w/w) fat – Standard chow (Harlan UK Ltd)
Gergerlioglu et al (2016) Expt 1 Wistar rats [Male] Formulated diet (Nukleon Ltd., Ankara, HFD 35% kcal from suet – Standard rat chow
Turkey)
Hargrave et al. (2016) Expt 1 Sprague Dawley rats Research Diets #D06040601 [6.1 kcal/g] Ketogenic Diet 80% kcal 15% Standard chow (Harlan Teklad)
[Male] [3.3 kcal/g]
Jamshed et al. (2014) Expt 1 Sprague Dawley rats Diets made from store bought ingredients by Cholesterol-Chocolate-Butterfat 5% (w/w) Butter-fat & 2% (w/ – Standard chow
[Male & Female] authors (CCB) diet w) Cholesterol

405
Jurdak et al (2008) Long–Evans rats [Male] Purina chow [3.6 kcal/g] with partially- HFD 9.0 kcal/g – Purina chow [3.6 kcal/g]
hydrogenated fat (traditional Crisco) [9.0 kcal/
g]
Liang et al. (2015) Sprague Dawley rats Formulated diet (Anlimo Technology, China) HFD 40% total kcal – Standard chow (Anlimo Technology)
[Male]
Magnusson et al (2015) C57BL/6 J mice [Male] Harlan Laboratories #TD.88137 [4.5 kcal/g] HFD 42% kcal 17.3% kcal PicoLab Rodent Diet #20 (LabDiet)
[4.07 kcal/g]
Mirzaei et al (2018) Wistar rats [Male] Royal Laboratory #D12492 [5.21 kcal/g] HFD 60% kcal 20% kcal Standard chow
Moreira et al (2014) Swiss Albino mice [Male] Chow is produced at the Universidade Estadual Hypercholestermic Diet 20% (w/w) Saturated fat + – Nuvilab-CR1 non-purified diet
de Campinas, Campinas, SP, Brazil 1.25% (w/w) Cholesterol
Pathan et al (2008) Sprague Dawley rats High fat pellet diet (Pranaw Agro Industries, HFD 58% total kcal 25% Standard chow diet (Pranaw Agro
[Male] New Delhi) Industries, New Delhi) [64 kcal/day]
Pratchayasakul et al (2015) Wistar rats [Male] Diet containing fat, mostly from lard HFD 59.28% total kcal (Lard) 26.45% C.P . Company #CP 082 Standard
[5.35 kcal/g] laboratory diet
Park et al (2018) ICR mice [Male] AIN-76A [3.77 kcal/g] + 40% beef Tallow No. HFD 40% (w/w) Beef Tallow – Standard Chow AIN-76A No. 100,000
101556 (Dyets Inc.) (Dyets Inc.) [3.77 kcal/g]
Scichilone et al (2016) Sprague Dawley rats Bio-Serv #F5155 supplemented with 14% Ketogenic Diet 73.5% (w/w) 14% Standard chow
[Male] protein
Silva et al. (2005) Wistar rats [Male] – Ketogenic Diet 69% (w/w) 24% Nuvilab-CR1non-purified diet
Spencer et al (2017) F344xBN F1 rats [Male] Teklands Diets #TD.06414 [5.1 kcal/g] HFD 60.3% total kcal 18% Teklad Diets Standard chow #TD. 8640
[3.0 kcal/g]
Spinelli et al (2017) C57BL/6 J mice [Male] Diets obtained from Mucedola (Italy) HFD 60% total kcal from saturated – Standard chow
fatty acids
Thirumangalakudi et al (2008) C57BL/6 mice Tekland Diets HFD 21% Fat + 1.25% Cholesterol – –
Valladolid-Acebes et al (2011) C57BL/6 J mice [Male] Research Diets #D12451. [4.73 kcal/g] HFD 45% kcal (20% w/w Lard) 20% kcal Research Diets Low-Fat Rodent Diet
#D12450B [3.85 kcal/g]
(continued on next page)
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
Table 2 (continued)

Author & Expt Species & Strain [sex] Experimental diet [Energy Density*] Experimental diet Label from Fat Content Protein Control Diet [Energy Density*]
Authors
K.N. Abbott, et al.

Valladolid-Acebes et al (2013) C57BL/6 J mice [Male] Research Diets #D12451. [4.73 kcal/g] HFD 45% kcal (20% w/w Lard) 20% kcal Standard chow
Expt 1
Valladolid-Acebes et al (2013) C57BL/6 J mice [Male] Research Diets #D12451. [4.73 kcal/g] HFD 45% kcal (20% w/w Lard) 20% kcal Standard chow
Expt 3
Wei et al. (2018) C57BL/6 J mice [Male] 74% control diet and 10% lard compound HFD 5.28% (w/w) in CD + 10% 5% (w/w) Regular Diet
(Shanghai Pu Lu Teng Biological Technology) (w/w) Lard
Woodie and Blythe (2018) Sprague Dawley rats Research Diets #D08060104 HFD 60% kcal – Harlan-Teklad Rat Chow #LM-485
[Male]
Xu et al. (2018) C57BL/6 J mice [Male] Research Diets #D12492 [5.21 kcal/g] HFD 60% kcal – Standard diet

406
Table 3
Composition and energy density of experimental and control diets provided to animals to examine the effect of high sugar diet (HSD) consumption on hippocampal-dependent learning and memory. (* = where
provided; w/w = weight / weight; kcal = kilocalories).
Author & Expt Species & Strain [sex] Experimental diet [Energy Density*] Experimental diet Label From Refined Sugar Content Protein Control Diet [Energy Density*]
Authors

Abbott et al (2016) Expt 1 Sprague Dawley rats [Male & Control diet [3.53 kcal/g] + 10% Sucrose HSD 0.4 kcal/g sucrose solution – Gordon’s Premium Rat and Mouse Breeder
Female] solution [0.4 kcal/g] diet (Australia) [3.53 kcal/g]
Beilharz et al (2014) Expt 2 Sprague Dawley rats [Male] Control diet [3.53 kcal/g] + 10% Sucrose HSD 0.4 kcal/g sucrose solution – Gordon’s Premium Rat and Mouse Breeder
solution [0.4 kcal/g] diet (Australia) [3.53 kcal/g]
Gergerlioglu et al (2016) Wistar rats [Male] Formulated diet (Nukleon Ltd., Ankara, Turkey) HSD 100% kcal of carbohydrate – Standard rat chow
Expt 1 from sucrose
Jurdak et al (2008) Long–Evans rats [Male] Purina chow [3.6 kcal/g] + 32% Sucrose solution HSD 1.28 kcal/g sucrose solution – Purina chow [3.6 kcal/g]
[1.28 kcal/g]
Kendig et al (2013) Hooded Wistar rats [Male & Specialty Feeds Standard Chow [3.4 kcal/g] + HSD 0.4 kcal/g sucrose solution – Specialty Feeds Standard chow [3.4 kcal/g]
Female] 10% sucrose solution [0.4 kcal/g]
Magnusson et al. (2015) C57BL/6 J mice [Male] Harlan Laboratories # TD.98090 [4 kcal/g] HSD 66% (w/w) sucrose 17.7% kcal PicoLab Rodent Diet #20 [4.07 kcal/g]
Xu and Reichelt (2018) Sprague Dawley rats [Male] Standard rat chow [2.63 kcal/g] + 10% sucrose HSD 0.4 kcal/g sucrose solution – Standard rat chow [2.63 kcal/g]
solution [0.4 kcal/g]
Woodie and Blythe (2018) Sprague-Dawley rats [Male] Research Diets #D05111802 High Fructose Diet 55% kcal – Standard chow
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
K.N. Abbott, et al.

Table 4
Composition and energy density of experimental and control diets provided to animals to examine the effect of high fat-high sugar diet (HFHS) consumption on hippocampal-dependent learning and memory. (* =
where provided; w/w = weight / weight; kcal = kilocalories).
Author & Expt Species & Strain Experimental diet [Energy Density*] Experimental diet Label Fat Content Refined Sugar Content Protein Control Diet [Energy Density*]
[sex] From Authors

Beilharz et al (2014) Expt 1 Sprague Dawley Cafeteria diet consisting of store-bought "junk Cafeteria Diet with 45% (w/w) 0.4 kcal/g Sucrose solution; 5% Gordon’s Premium Rat and Mouse
rats [Male] foods" + 10% Sucrose Solution [0.4 kcal/g] Sucrose solution Unknown from solid diet Breeder diet (Australia) [3.53 kcal/g]
Beilharz et al (2014) Expt 2 Sprague Dawley Cafeteria diet consisting of store-bought "junk Cafeteria Diet with 45% (w/w) 0.4 kcal/g Sucrose solution; 5% Gordon’s Premium Rat and Mouse
rats [Male] foods" + 10% Sucrose Solution [0.4 kcal/g] Sucrose solution Unknown from solid diet Breeder diet (Australia) [3.53 kcal/g]
Cafeteria diet consisting of store-bought "junk Cafeteria Diet 45% (w/w) Unknown 5%
foods"
Beilharz et al (2018) Sprague Dawley Cafeteria diet consisting of store-bought "junk Cafeteria Diet with 45% (w/w) 0.4 kcal/g Sucrose solution; 5% Standard chow
rats [Male] foods" + 10% Sucrose Solution [0.4 kcal/g] Sucrose solution Unknown from solid diet
Hargrave et al (2016) Expt Sprague Dawley Harlan #TD.10768 [4.5 kcal/g] High-fat, high-dextrose 38% kcal 20% kcal Dextrose 24% Harlan Teklad Standard chow [3.3 kcal/
1 rats [Male] “Western” diet g]

407
Kanoski and Davidson Sprague Dawley Harlan Teklad Powered Diet #TD.04489 High Energy Diet 170 g/kg Lard 220.5 g/kg Glucose – Purina LabDiet #5001 [3.0 kcal/g]
(2010) rats [Male] [4.55 kcal/g]
Molteni et al (2004) Fisher 344 rats Powdered HFS diet (Purina Mills Inc., Test Diets HFHS 39% total kcal 40% total kcal from sucrose – Powdered low-fat, complex
[Female] Inc.) (primarily lard) carbohydrate diet (Purina Mills Inc.,
Test Diets Inc.)
Ross et al (2013) Sprague Dawley Control diet, a petri dish containing animal lard HFHS Free access to Lard 32% Sucrose solution; – Purina LabDiet #5001 [3.0 kcal/g]
rats [Male] [9.0 kcal/g], and a bottle of 32% sucrose solution Unknown from solid diet
[1.13 kcal/g]
Takechi et al (2017) C57BL6/J mice Specialty Feeds #SF14-088 High Fat Fructose Diet 30% (w/w) Lard, 0.5% 15% (w/w) Fructose – Specialty Feeds Low-Fat chow #AIN-
[Male] (w/w) Cholesterol 93M
Tran and Westbrook Sprague Dawley Cafeteria diet made of store-bought "junk foods" HFHS Diet 40% (w/w) Primarily 40% (w/w) Sucrose 5% Gordon’s Premium Rat and Mouse
(2015) Expt 1 rats [Male] lard Breeder diet (Australia) [3.53 kcal/g]
Tran and Westbrook Sprague Dawley Cafeteria diet made of store-bought "junk foods" HFHS Diet 40% (w/w) Primarily 40% (w/w) Sucrose 5% Gordon’s Premium Rat and Mouse
(2018) Expt 1 rats [Male] lard Breeder diet (Australia) [3.53 kcal/g]
Wu et al (2003) Expt 1 Sprague Dawley Powdered HFHS diet (Purina Mills Inc., Test Diets HFS Diet 39% total kcal 40% total kcal from sucrose – Powdered low-fat, complex
rats [Male] Inc.) (primarily lard) carbohydrate diet (Test Diets Inc.)
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
Table 5
Experimental details and results for studies examining the effect of dietary manipulation on hippocampal-dependent learning and memory in the Morris water maze. (*Approximate age based on growth charts. When
multiple groups of diet manipulation lengths have been used: + = separate experimental groups; When multiple tests have been used: ^ = ongoing test without training between test sessions; # = isolated test involving
training and testing. CD = Control Diet; HFD = high fat diet; HFHS = high fat-high sugar diet; HSD = high sugar diet).
K.N. Abbott, et al.

Author & Expt Species & Strain [sex] Experiment Diet Age* at Diet Onset Diet Dependent Variable Results
Duration [Retention Interval]

Boitard et al (2014) Wistar rats [Male] HFD 12 weeks 2 months Number target crossings [2- Rats consuming HFD had an equivalent number of target crossings as
hr]^ control rats at both 2-hr and 4-day retention intervals.
Number target crossings [4
day]^
Che et al (2018) SAMP8 mice [Male] HFD 10 months old 8 weeks Time in target quadrant [24- Mice consuming HFD spent significantly less time in the target quadrant
hr] than control mice
Denver et al (2018) C57BL/6 J mice [Male] HFD 7-14 weeks 10 days+ Time in target quadrant [24- Rats consuming the HFD did not significantly differ from rats consuming
26 days+ hr] CD, however the time in target quadrant for both groups did not
8 weeks+ significantly differ from chance.
Gergerlioglu et al (2016) Wistar rats [Male] HFD 10–12 weeks 3 weeks Time in target quadrant [24- Rats consuming HFD spent significantly less time in the target quadrant
Expt 1 hr] than rats consuming CD.
HSD 10–12 weeks 3 weeks Time in target quadrant [24- There was no effect of diet on time spent in the target quadrant.
hr]
Jamshed et al (2014) Expt Sprague Dawley rats HFD (labelled cholesterol- 6+ weeks* (180- 6 weeks Escape latency [24-hr] Rats consuming the CCB diet took significantly longer to reach the escape
1 [Male & Female] chocolate-butterfat diet by 200 grams) platform than rats consuming CD.
authors)
Jurdak et al (2008) Long–Evans rats [Male] HFD 6 weeks 5 weeks Time in target quadrant [24- There was no effect of diet on time spent in the target quadrant at 24-hrs or
hr]^ 10-days post-training.
Time in target quadrant [10
day]^
HSD 6 weeks 5 weeks Time in target quadrant [24- Rats consuming the HSD spent equivalent time in the target quadrant as
hr]^ rats consuming CD when tested 24-hrs after training, however spent

408
Time in target quadrant [10 significantly less time in the target quadrant than control rats when tested
day]^ 10-days after training.
Kendig et al (2013) Hooded Wistar rats [Male HSD 4 and 8 weeks 4 weeks Time in target quadrant During the first three probe trials, rats consuming HSD spent significantly
& Female] (Probe tests 1 -3) [24-hr]^ less time in the target quadrant than rats consuming CD.
Time in target quadrant [28 Rats consuming HSD spent less time in the target quadrant than rats
day]^ consuming CD, however this difference was not significant.
Liang et al. (2015) Sprague Dawley rats HFD 6 weeks 6 weeks Time in target quadrant [24- Rats consuming HFD spent significantly less time in the target quadrant
[Male] hr] than rats consuming CD.
Magnusson et al. (2015) C57BL/6 J mice [Male] HFD 8 weeks 5 weeks Proximity from platform Rats consuming HFD had greater proximity from platform location than
(Probe 1) [1-hr]# control rats, however this difference was not significant.
Proximity from platform Rats consuming HFD had greater proximity from platform location than
(Probe 2) [1-hr]# control rats, however this difference was not significant.
Proximity from platform Rats consuming HFD had greater proximity from platform location than
(Probe 3) [1-hr]# control rats, however this difference was not significant.
Proximity from platform Rats consuming HFD had a smaller proximity from platform location than
(Probe 4) [1-hr]# rats consuming chow, however this was not significant.
Reversal Task [1-hr] Proximity from new platform location did not differ between HFD and CD
rats; HFD had significantly smaller proximity to old platform location than
CD rats.
HSD Proximity from platform Rats consuming HSD had greater proximity from platform location than
(Probe 1) [1-hr]# control rats, however this was not significant.
Proximity from platform No effect of diet on proximity from platform, with rats in both groups
(Probe 2) [1-hr]# having a similar distance from the platform location.
Proximity from platform Rats consuming HSD had a smaller proximity from platform location than
(Probe 3) [1-hr]# control rats, however this was not significant.
Proximity from platform Rats consuming HSD had a smaller proximity from platform location than
(Probe 4) [1-hr]# control rats, however this was not significant.
(continued on next page)
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
Table 5 (continued)

Author & Expt Species & Strain [sex] Experiment Diet Age* at Diet Onset Diet Dependent Variable Results
Duration [Retention Interval]
K.N. Abbott, et al.

Reversal Task [1-hr] Proximity from new platform location did not differ between HSD and CD
rats; HSD had significantly smaller proximity to old platform location than
CD rats.
Mirzaei et al (2018) Wistar rats [Male] HFD 8 Weeks 8 weeks Time in target quadrant [24- Rats consuming a HFD spent significantly less time in the target quadrant
hr] than rats consuming CD.
Molteni et al (2004) Fisher 344 rats [Female] Powdered HFHS 8 weeks 2 months Time in target quadrant [3 Rats consuming HFHS diet spent equal time in all four MWM quadrants at
day] test; time in target quadrant was significantly lower than that of CD rats.
Pathan et al (2008) Sprague Dawley rats HFD 5+ weeks* (150- 5 weeks Escape latency [24-hr] Decrease in escape latency across the five days of training was significantly
[Male] 190 grams) greater in rats consuming CD than rats consuming HFD.
Pratchayasakul et al (2015) Wistar rats [Female] HFD 8+ weeks* (200- 4 weeks+ Time in target quadrant [24- There was no effect of diet on time in target quadrant.
220 grams) hr]
8 weeks+ Time in target quadrant [24- There was no effect of diet on time in target quadrant.
hr]
Scichilone et al (2016) Sprague Dawley rats Ketogenic Diet 4 weeks 8 weeks Time in target quadrant [2- Rats consuming the ketogenic diet spent less time in the target quadrant
[Male] hr] than control rats, however this was not significant.
Silva et al (2005) Wistar rats [Male] Ketogenic Diet 30 days old 40 days Time in target quadrant [24- Rats consuming the ketogenic diet spent less time in the target quadrant
hr] than control rats, however this was not significant.
Spencer et al (2017) F344xBN F1 rats [Male] HFD 3-months+ 4 days Time in target quadrant [4 There was no effect of diet on time spent in the target quadrant.
24-months+ day] There was no effect of diet on time spent in the target quadrant.
Spinelli et al (2017) C57BL/6 J mice [Male] HFD 30-35 days 6 weeks Time in target quadrant [24- Mice consuming HFD spent significantly less time in the target quadrant
hr] than control mice.
Takechi et al (2017) C57BL6/J mice [Male] High Fat Fructose Diet (or 6 weeks 4 weeks Escape latency [24 -hr] Mice consuming HFF diet had a slower latency to the platform than mice
HFHS) consuming CD, however this was not significant.
Wei et al (2018) C57BL/6 J mice [Male] HFD 12-14 months 8 Weeks Time in target quadrant [24- HFD mice spent less time in the target quadrant than mice consuming CD.
hr]

409
Woodie and Blythe (2018) Sprague Dawley rats HFD 6 weeks 8 weeks Time in target quadrant [24- There was no effect of diet on the time spent in the target quadrant during
[Male] hr] probe test.
Reversal Task Time in target There was no effect of diet on time spent in new target quadrant.
quadrant [24-hr]
High Fructose Diet (or HSD) Time in target quadrant [24- There was no effect of diet on the time spent in the target quadrant during
hr] probe test.
Reversal Task Time in target There was no effect of diet on time spent in new target quadrant.
quadrant [24-hr]
Wu et al (2003) Expt 1 Sprague Dawley rats Powdered HFHS 6+ weeks* (200- 4 weeks Time in target quadrant [4- There was no effect of diet on time spent in the target quadrant.
[Male] 240 grams) hr]
Xu et al (2018) C57BL/6 J mice [Male] HFD 4 Weeks 8 weeks Time in target quadrant [24- HFD spent significantly less time in the target quadrant than mice
hr] consuming CD.
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

experiments (some studies conducted multiple tests within a single 3.2.5. Place recognition
experiment) for inclusion in the analysis. Experimental animals were A total of 391–394 (control n = 189–190; experimental diet
provided ad libitum access to a formulated high fat diet obtained from a n = 202–204) animals were used to assess the effect of dietary intake
local manufacturer, except for Jamshed et al. (2014) who used a diet on spatial memory using the place recognition task. Of the 13 experi-
consisting of store-bought foods, and Jurdak et al. (2008) and Wei et al. ments, four examined the effect of a high sugar diet provided ad libitum
(2018) who used control chow mixed with partially-hydrogenated fat (Beilharz et al., 2014; Magnusson et al., 2015) or 2 -hour restricted
or lard, respectively. access (Abbott et al., 2016; Xu and Reichelt, 2018), five examined the
effect of high fat diet (Ayabe et al., 2018; Magnusson et al., 2015;
3.2.2. High sugar diet Moreira et al., 2014; Valladolid-Acebes et al., 2011, 2013), and five
A total of 241 rats and mice (control n = 117; high sugar diet examined a high fat-high sugar diet (Beilharz et al., 2014, 2018; Ross
n = 124) were used to assess the effect of a high sugar diet intake. Of et al., 2013; Tran and Westbrook, 2015, 2018). All experiments used
the nine experiments, four used Sprague-Dawley rats (Abbott et al., male subjects, with the exception of Abbott et al. (2016) who used male
2016; Beilharz et al., 2014; Kendig et al., 2013; Xu and Reichelt, 2018), and female rats.
two used C57BL6 J mice (Magnusson et al., 2015), one used Wistar rats Twenty-one dependent variables were extracted from the 13 ex-
(Gergerlioglu et al., 2016), one used hooded Wistar rats (Kendig et al., periments for inclusion in the analysis. All of the experiments assessed
2013), and one used Long Evans rats (Jurdak et al., 2008). All experi- spatial memory using a score to quantify the difference in time spent
ments used male subjects, with the exception of Abbott et al. (2016) exploring the object in a novel location compared with the time spend
and Kendig et al. (2013) who used male and female rats. exploring the object in a familiar location. The exact formula for cal-
A total of 18 dependent variables were extracted from the nine culating the memory score sometimes differed between studies (e.g.,
experiments for inclusion in the analysis. Experimental animals were novelty proportion vs discrimination ratio), but more time spent ex-
provided a 10% sucrose solution ad libitum (Beilharz et al., 2014; ploring the novel location was always taken to index better recognition
Jurdak et al., 2008) or in a 2 -hour daily access period (Abbott et al., memory.
2016; Kendig et al., 2013; Xu and Reichelt, 2018), a 32% sucrose so-
lution with ad libitum access (Jurdak et al., 2008), or a semi-purified 3.2.6. Radial arm or radial arm water maze
rodent chow with a high sugar content (Gergerlioglu et al., 2016; A total of 147–153 (control n = 72–75; experimental diet
Magnusson et al., 2015; Woodie and Blythe, 2018). n = 75–78) animals were used to assess the effect of dietary intake on
spatial memory using either the radial arm or radial arm water-maze
task. Of the six experiments, five examined the effect of high fat diet
3.2.3. High fat-high sugar diet intake (Alzoubi et al., 2013a, 2013b, 2018; Thirumangalakudi et al.,
A total of 303 rats and mice (control n = 155; high fat-high sugar 2008; Valladolid-Acebes et al., 2011) and one examined the effect of
diet n = 148) were used to assess the effect of a high fat-high sugar diet. high fat-high sugar diet intake (Kanoski and Davidson, 2010). All ex-
Of the 11 experiments, nine used Sprague-Dawley rats (Beilharz et al., periments used male subjects.
2014, 2018; Hargrave et al., 2016; Kanoski and Davidson, 2010; Ross Thirteen dependent variables were extracted from the six experi-
et al., 2013; Tran and Westbrook, 2015, 2018; Wu et al., 2003), one ments for inclusion in the analysis. All experiments assessed spatial
used C57BL/6 J mice (Takechi et al., 2017), and one used F344 rats memory by measuring the total sum of working memory errors, defined
(Molteni et al., 2004). All experiments used male subjects, with the as re-entry into an arm that previously contained a reward or platform,
exception of Molteni et al (2004) who used female rats. and reference memory errors, defined as entry or re-entry an arm that
A total of 13 dependent variables were extracted from 11 experi- has never contained a reward or platform; two experiments evaluated
ments for inclusion in the analysis. Experimental animals were pro- total errors in addition to working memory errors separately
vided ad libitum access to a formulated high fat-high sugar diet in (Thirumanagalakudi et al., 2008; Valladolid-Acebes et al., 2011).
pelleted (Hargrave et al., 2016; Takechi et al., 2017) or powdered
(Kanoski and Davidson, 2010; Molteni et al., 2004; Wu et al., 2003) 3.2.7. Spontaneous alternation
form, a diet consisting of chow supplemented with store-bought foods A total of 76 (control n = 43; experimental diet n = 33) animals
(Beilharz et al., 2014, 2018; Tran and Westbrook, 2015, 2018), or chow were used to assess the effect of diet on spatial memory using sponta-
supplemented with lard and a bottle of 32% sucrose solution (Ross neous alternation in either a T- (Arnold et al., 2014) or Y-maze
et al., 2013). (Hargrave et al., 2016; Park et al., 2018). All four experiments ex-
amined the effect of a high fat diet (Arnold et al., 2014; Hargrave et al.,
3.2.4. Morris water maze 2016; Park et al., 2018), and one additionally examined the effect of
A total of 531–544 (control n = 258–266; experimental diet high fat-high sugar diet (Hargrave et al., 2016). All experiments used
n = 273–278) animals were used to assess the effect of dietary intake male subjects.
on spatial memory using the Morris water maze. Of the 22 experiments, A total of seven dependent variables were extracted from the four
five examined the effect of a high sugar diet (Gergerlioglu et al., 2016; experiments for inclusion in the analysis. One experiment assessed
Jurdak et al., 2008; Kendig et al., 2013; Magnusson et al., 2015; Woodie spatial memory by measuring the percentage of sequential arm entries
& Blythe, 2018), 18 examined the effect of high fat diet, and three (Hargrave et al., 2016), and the remaining three examined the number
examined the effect of high fat-high sugar diet (Molteni et al., 2004; of entries into a new arm in the T- or Y-maze.
Takechi et al., 2017; Wu et al., 2003). All experiments used male sub-
jects, except for Jamshed et al. (2014) and Kendig et al. (2013) who 3.3. Meta-analysis output
used male and female rats, and Molteni et al. (2004) and
Pratchayasakul et al. (2015) who used female rats. The forest plots and individual statistics for the diets high in fat,
Thirty-nine dependent variables were extracted from the 22 ex- sugar, and both fat and sugar are shown in Figs. 2–4, respectively.
periments for inclusion in the analysis. All of the experiments used time There was a significant overall negative effect of consuming a diet high
in target quadrant as their dependent variable of spatial memory, ex- in fat (k = 53, g = -0.595, 95% CI [-0.715 – -0.474], p < .001), re-
cept for those reported by Boitard et al. (2014), which used number of fined sugar (k = 18, g = -0.552, 95% CI [-0.750 – -0.355], p < .001),
target crossings, Magnusson et al. (2015), which used proximity from or both fat and refined sugar (k = 13, g = -0.654, 95% CI [-0.882 –
platform location, and, finally, by Jamshed et al. (2014); Pathan et al -0.426], p < .001), on hippocampal-dependent spatial learning and
(2008), and Wei et al (2018) which all used escape latency. memory. Each diet had a medium effect size.

410
Table 6
Experimental details and results for studies examining the effect of dietary manipulation on hippocampal-dependent learning and memory in the place recognition task. (*Approximate age based on growth charts. When
multiple groups of diet manipulation lengths have been used: + = separate experimental groups; When multiple tests have been used: ^ = ongoing test without training between test sessions; # = isolated test involving
training and testing. CD = control diet; HFD = high fat diet; HFHS = high fat-high sugar diet; HSD = high sugar diet).
Author & Expt Species & Strain Experiment Diet Age* at Diet Onset Diet Behavioural Task Results
K.N. Abbott, et al.

(sex) Duration [Retention Interval]

Abbott et al (2016) Expt 1 Sprague Dawley rats HSD 4 weeks 2 weeks Place Recognition [5- Male and female HSD rats spent a significantly smaller proportion of time
[Male & Female] min] exploring the novel object location than control rats.
3 weeks Object-in-Place Male HSD rats spent a significantly smaller proportion of time exploring the
Recognition [5-min] objects in the novel locations than male control rats. There was no effect of
HSD on exploration of objects in female rats.
Ayabe et al (2018) C57BL/6 J mice HFD 6 Weeks 7 Weeks Place Recognition [4- HFD rats spent a smaller proportion of time exploring the novel object
[Male] hr] location then control mice, however this was not significant.
Beilharz et al (2014) Expt 1 Sprague Dawley rats HFHS (labelled Cafeteria Diet with 8+ weeks* (363- 5-6 days Place Recognition [5- Main effect of diet, such that rats consuming cafeteria diet with sucrose
[Male] Sucrose solution by authors) 463 grams) min]# solution spent a significantly smaller proportion of time exploring the novel
11-12 days Place Recognition [5- object location than control rats.
min]#
20 - 21 days Place Recognition [5-
min]#
Beilharz et al (2014) Expt 2 Sprague Dawley rats HSD 7+ weeks* 5-6 days Place Recognition [5- Main effect of diet, such that rats consuming the HSD, cafeteria diet, or the
[Male] (302–366 grams) min]# cafeteria diet with sucrose solution spent a significantly smaller proportion of
HFHS (labelled Cafeteria Diet with 11-12 days Place Recognition [5- time exploring the novel object location than control rats.
Sucrose solution by authors) min]#
HFHS (labelled Cafeteria Diet with 20 - 21 days Place Recognition [5-
Sucrose solution by authors) min]#
Beilharz et al (2018) Sprague Dawley rats HFHS (labelled Cafeteria Diet with 5+ weeks* (200 grams) 10-11 days Place Recognition [5- Main effect of diet, such that rats consuming cafeteria diet with sucrose
[Male] Sucrose solution by authors) min]# solution spent a significantly smaller proportion of time exploring the novel
17-18 days Place Recognition [5- object location than control rats.
min]#

411
Magnusson et al (2015) C57BL/6 J mice HFD 8 weeks 2 weeks Place Recognition [1- There was no significant effect of diet on proportion of time spent exploring
[Male] hr]^ the object in the novel object location, however exploration did not occur
Place Recognition [24- above chance for either HFD or CD animals.
hr]^
HSD 8 Weeks 2 Weeks Place Recognition [1- There was no significant effect of diet on proportion of time spent exploring
hr]^ the object in the novel object location, however exploration did not occur
Place Recognition [24- above chance for either HSD or CD animals.
hr]^
Moreira et al (2014) Swiss Albino mice HFD (labelled Hypercholestermic Diet 3 months old 8 weeks Place Recognition [90- Mice consuming a hypercholesterolemic diet spent a significantly smaller
[Male] by authors) min] proportion of time exploring the novel object location than control mice.
Ross et al (2013) Sprague Dawley rats HFHS 7-8 weeks 4 weeks Place Recognition [24- There was no effect of diet on the proportion of time spent exploring the
[Male] hr] novel object location.
Tran and Westbrook (2015) Sprague Dawley rats HFHS 5+ weeks* (200- 1 week Place Recognition [5- Main effect of diet, such that rats consuming HFHS diet spent a significantly
Expt 1 [Male] 300 grams) min]# smaller proportion of time exploring the novel object location than control
2 week Place Recognition [5- rats.
min]#
3 week Place Recognition [5-
min]#
Tran and Westbrook (2018) Sprague Dawley rats HFHS 5+ weeks* (200- 1 week Place Recognition [5- Rats consuming HFHS diet spent a significantly smaller proportion of time
Expt 1 [Male] 300 grams) min] exploring the novel object location than control rats
Valladolid-Acebes et al C57BL/6 J mice HFD 5 weeks 8 weeks Place Recognition [1- Mice consuming HFD spent a significantly smaller proportion of time
(2013) Expt 1 [Male] hr]^ exploring the novel object location than control mice tested 1-hr and 24 -hrs
Place Recognition [24- post-training.
hr]^
Valladolid-Acebes et al C57BL/6 J mice HFD 8 weeks 8 weeks Place Recognition [1- There was no effect of diet on proportion of time spent exploring the novel
(2013) Expt 3 [Male] hr]^ object location when tested 1-hr following training; HFHS diet mice spent a
Place Recognition [24- smaller proportion of time exploring the novel object location at 24-hr test,
hr]^ however this difference was not significant.
Xu and Reichelt (2018) Sprague Dawley rats HSD 4 weeks 3 weeks Place Recognition [5- HSD rats spent a significantly smaller proportion of time exploring the novel
[Male] min] object location than control rats.
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

The forest plots and individual statistics for the Morris water maze, papers (Alzoubi et al., 2013a; and 2013b), which may have biased the
place recognition task, radial arm mazes, and spontaneous alternation results. However, the weighted effect sizes of the six dependent mea-
tasks are shown in Figs. 5–8, respectively. There was a significant sures extracted from these two papers were among the smaller effect
overall negative effect of consuming a diet high in fat, sugar, or high in sizes in the analysis, suggesting that the large SMD observed for this
both fat and sugar, on performance in the Morris water maze (k = 43, g task group does not rely on their inclusion in the analysis.
= -0.375, 95% CI [-0.511 – -0.239], p < .001), place recognition task The role of the hippocampus in successful performance on spatial
(k = 21, g = -0.759, 95% CI [-0.939 – -0.578], p < .001), radial arm learning and memory tasks is well established (Barker and Warburton,
or radial arm water maze (k = 13, g = -0.932, 95% CI [-1.153 – 2011; Compton et al., 1997; Floresco et al., 1997; Lalonde, 2002), but
-0.710], p < .001), and spontaneous alternation (k = 7, g = -0.557, studies differed in the procedures used in these tasks. Moreover, var-
95% CI [-0.894 – -0.219], p < .01). The effect of diet on behavioral iations in how these tasks are implemented may affect the degree to
performance was a medium effect size for all tasks except for the radial which the hippocampus is recruited. For example, studies that used the
arm maze where the effect size was large. Morris water maze differed in terms of the number of training days, the
number of training trials per day, the inter-trial and inter-training day
4. Discussion intervals, and the retention interval before test. Similarly, there were
differences across studies in the place recognition task, such as the
There is considerable evidence that rodents fed a diet high in fat, number of objects used, the length of familiarization time, the retention
sugar, or both fat and sugar exhibit impairments in tasks assessing interval before test, and the arena where the task was conducted.
hippocampal-dependent forms of learning and memory (Kanoski et al., Procedural variations in these tasks, particularly in the amount of
2010; Yeomans, 2017). However, there have been failures to detect training, length of retention interval, and maze set-up (affecting the
such impairments (e.g., Ayabe et al., 2018; Ross et al., 2013, Jurdak amount of extra-maze spatial information that can be extracted), have
et al., 2008). One source of the inconsistent results is the type of diet been argued to affect the involvement of brain regions outside the
used, while a second source is the type of task used to assess the effects hippocampus (e.g., Tran and Westbrook, 2017; Vorhees and Williams,
of the diet. Therefore, we subjected this evidence to a meta-analysis in 2006, 2014; see Tran and Westbrook, 2018 for discussion in relation to
order to provide a quantitative assessment of the protocols under which object recognition memory). Thus, such variations may have affected
dietary manipulations adversely affect hippocampal-dependent forms the difficulty of the task; increasing or decreasing the reliance on a
of cognition. We had two aims: The first was to assess the contribution functioning hippocampus for successful performance. Moreover, the
of diets high in saturated fat, refined sugar, or both, to impairments of effects of diet on learning and memory can be highly specific, impairing
spatial learning and memory. The second aim was to examine the tasks some, but not other, similar tasks (e.g., Hargrave et al., 2016; Tran and
used to determine whether they differed in their sensitivity to the Westbrook, 2015). Therefore, the inconsistencies in the literature on
dietary manipulations. the observed relationship between diet and cognition are likely a con-
sequence of procedural difference in the assessment measure rather
4.1. Summary of results than the content of dietary manipulation.

The meta-analysis revealed that all three diet manipulations pro- 4.2. Rodent species as a variable for diet-induced cognitive impairment
duce a significant impairment in spatial learning and memory tasks,
relative to a control diet that is usually starch based and low in satu- Rats and mice, as well as different strains within these species, can
rated fats and refined sugars. While direct statistical comparisons be- perform differently on hippocampal-dependent behavioral tasks (e.g.
tween the diet manipulation categories were not permitted, the size of Cressant et al., 2007; Ennaceur et al., 2005; Genzel et al., 2017;
the Hedge’s g for the three manipulations suggests that the largest effect Stranahan, 2011; reviewed in Hok et al., 2016; Keeley et al., 2015).
size was found for studies that used a diet high in both fat and sugar, Such differences are thought to be due to variations in metabolism
followed closely by studies using a high fat diet, and, finally, a high (Chang et al., 1990), hippocampal neurogenesis (Merritt and Rhodes,
sugar diet. Nonetheless, all three dietary manipulations were observed 2015), gene expression (Malki et al., 2015), brain volume (Keeley et al.,
to have a medium effect size, which suggests that differences in dietary 2015), and place cell activity (Ji et al., 2018; see Manahan-Vaughan,
composition are unlikely to be the single or primary factor responsible 2019). These differences in turn could interact with dietary manipula-
for the inconsistent findings in the literature. tions to affect performance on hippocampal-dependent tasks (Buettner
A second source of variability in the findings reported is the beha- et al., 2007; Stöckli et al., 2017; West et al., 1992, 1995). However, we
vioral task used to assess dietary effects on learning and memory. As found little if any evidence for such effects based on the study de-
noted in the Introduction, tasks assessing hippocampal forms of cog- scriptives.
nition, specifically, spatial learning and memory, appear to be espe- Of the 41 articles that were included in the present meta-analysis,
cially sensitive to dietary manipulations. The role of the hippocampus 16 experiments were extracted from 14 articles examining dietary
in successful performance on the tasks selected for inclusion in the manipulations in mice and 29 experiments were extracted from 27 ar-
meta-analysis has been well documented: the Morris water maze (e.g., ticles examining dietary manipulations in rats. Impairment in hippo-
Compton et al., 1997), place recognition (e.g., Barker and Warburton, campal-dependent learning and memory was observed in 10 experi-
2011), radial arm and radial arm water maze (e.g., Floresco et al., ments (approximately 63%) using mice (Arnold et al., 2014; Che et al.,
1997), spontaneous alternation (e.g., Lalonde, 2002). When the per- 2018; Moreira et al., 2014; Park et al., 2018; Spinelli et al., 2017;
formance of animals consuming a diet high in saturated fat, refined Thirumangalakudi et al., 2008; Valladolid-Acebes et al., 2011, 2013,
sugar, or both, on these tasks were considered separately, the meta- Experiment 1; Wei et al., 2018; Xu et al., 2018) and 16 experiments
analysis revealed a significant diet-induced impairment in spatial (approximately 55%) using rats (Abbott et al., 2016; Alzoubi et al.,
learning and memory for each task group. Again, direct statistical 2013a, 2013b, 2018; Beilharz et al., 2014; Kanoski and Davidson, 2010;
comparisons between the behavioral tasks was not permitted. However, Liang et al., 2015; Mirzaei et al., 2018; Molteni et al., 2004; Pathan
the largest impairment was seen on the radial arm and radial water et al., 2008; Tran and Westbrook, 2015, 2018; Xu and Reichelt, 2018);
maze tasks, where there was a large effect size, followed by the place no deficits were observed in five experiments using mice (Ayabe et al.,
recognition task, the spontaneous alternation task, and the Morris water 2018; Denver et al., 2018; Magnusson et al., 2015; Takechi et al., 2017)
maze task, each of which had a medium effect size. It should be noted and nine experiments using rats (Boitard et al., 2014; Jamshed et al.,
that nearly half of the experiments included in the meta-analysis for the 2014; Pratchayasakul et al., 2015; Ross et al., 2013; Scichilone et al.,
radial arm maze and radial water maze tasks were extracted from two 2016; Silva et al., 2005; Spencer et al., 2017; Woodie and Blythe, 2018;

412
K.N. Abbott, et al.

Table 7
Experimental details and results for studies examining the effect of dietary manipulation on hippocampal-dependent learning and memory in the radial arm maze (RAM) or radial arm water maze (RAWM).
(*Approximate age based on growth charts. When multiple groups of diet manipulation lengths have been used: + = separate experimental groups; When multiple tests have been used: ^ = ongoing test without training
between test sessions; # = isolated test involving training and testing. CD = control diet; HFD = high fat diet; HFHS = high fat-high sugar diet; HSD = high sugar diet; RM = reference memory; WM = working
memory;).
Author & Expt Species & Strain Experimental Diet Age* at Diet Onset Diet duration Behavioural Task (Dependent Results
[sex] Variable) [Retention Interval]

Alzoubi et al (2013a) Expt 1 Wistar rats HFD (labelled Western Diet by 6-7 weeks 6 weeks Radial 6-arm water maze (WM & Rats consuming the western diet made significantly more incorrect
[Male] authors) RM errors) [30-min]^ arm entries than rats consuming control diet when tested 30-min, 5 -
Radial 6-arm water maze (WM & hrs, and 24 -hrs post-training
RM errors) [5-hr]^
Radial 6-arm water maze (WM &
RM errors) [24-hr]^
Alzoubi et al (2013b) Expt 1 Wistar rats HFD (labelled High Fat 6-7 weeks 6 weeks Radial 6-arm water maze (WM & Rats consuming the HFCD made significantly more incorrect arm
[Male] Cholesterol Diet by authors) RM errors) [30-min]^ entries than rats consuming control diet when tested 30-min, 5 -hrs,
Radial 6-arm water maze (WM & and 24 -hrs post-training.

413
RM errors) [5-hr]^
Radial 6-arm water maze (WM &
RM errors) [24-hr]^
Alzoubi et al (2018) Wistar rats HFD 4+ weeks* 4 weeks Radial 6-arm water maze (WM & Rats consuming the HFD made significantly more incorrect arm
[Male] (160–200 grams) RM errors) [5-hr]^ entries than rats consuming control diet when tested 5 -hrs and 24 -
Radial 6-arm water maze (WM & hrs post-training.
RM errors) [24-hr]^
Kanoski and Davidson (2010) Sprague Dawley HFHS (labelled High Energy 8-9 weeks 6 tests starting Radial Arm Maze (WM & RM Rats consuming the high energy diet made significantly more
rats [Male] Diet by authors) at 3 days errors) [training prior to diet incorrect arm entries than control rats.
access]
Thirumangalakudi et al C57BL/6 mice HFD 4 months old 2 months Radial 8-Arm Water Maze (WM Mice consuming HFD made significantly more entries into an arm
(2008) Errors) [24-hrs] that previously contained a platform than mice consuming CD.
Radial 8-Arm Water Maze (WM & Mice consuming the HFD made significantly more incorrect arm
RM errors) [same test] entries than mice consuming tCD.
Valladolid-Acebes et al. C57BL/6 J mice HFD 4 weeks old 8 weeks Radial arm maze (WM errors) [24- Mice consuming HFD made significantly more entries into a
(2011) [Male] hrs] previously baited arm than mice consuming CD.
Radial arm maze (WM & RM Mice consuming the HFD made significantly more incorrect arm
errors) [same test] entries than mice consuming CD.
Neuroscience and Biobehavioral Reviews 107 (2019) 399–421
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

on growth charts. When multiple groups of diet manipulation lengths have been used: + = separate experimental groups; When multiple tests have been used: ^ = ongoing test without training between test sessions; # =
Experimental details and behavioural outcomes for studies examining the effect of dietary manipulation on hippocampal-dependent learning and memory in the spontaneous alternation task. (*Approximate age based
Wu et al., 2003). Mixed results were reported in one study using mice,

Rats consuming the western diet had a significantly lower


Mice consuming the HFD t showed significantly reduced

There was no effect of diet on percentage of sequential

There was no effect of diet on percentage of sequential

There was no effect of diet on percentage of sequential


in which a deficit was detected on a place recognition task after a 24 -

Mice consuming the HFD showed significantly reduced

Mice consuming the HFD showed significantly reduced


percentage of sequential arm entries than control rats.
hour retention period but not a 1 -hour retention period (Valladolid-

spontaneous alternation than mice consuming CD.


spontaneous alternation than rats consuming CD.

spontaneous alternation than rats consuming CD.


Acebes et al., 2013, Experiment 3). Mixed results were also reported in
four experiments using rats: one study reported a deficit on time in
target quadrant on the Morris water maze in rats consuming HFD but
not HSD (Gergerlioglu et al., 2016); one reported a deficit on the Y-
maze after 10 days of a high fat-high sugar diet, but no deficit after 40
days of the diet, nor after 10 or 40 days of a high fat diet (Hargrave
et al., 2016); one reported a deficit measuring time spent in target
Behavioural Outcomes

quadrant on the Morris water maze task at a 10 day retention interval


but not a 24-hour retention interval (Jurdak et al., 2008), and one re-
ported impaired learning in the Morris water maze task but no differ-
arm entries.

arm entries.

arm entries.
ence in the time spent in the target quadrant at test (Kendig et al.,
2013). Thus, diet-induced impairments on hippocampal-dependent
spatial learning and memory do not appear to be specific to rodent
species.
Behavioural Task (Dependent Variable)

(percentage sequential arm entries)

4.3. Sex as a variable for diet-induced cognitive impairment


Y-Maze Spontaneous Alternation
T-Maze Spontaneous Alternation

T-Maze Spontaneous Alternation

Y-maze Spontaneous Alternation

Sex differences in hippocampal-dependent spatial learning and


memory are well documented in both humans and rodents (reviewed in
Keeley et al., 2015; Koss and Frick, 2017; Yagi and Galea, 2018), with
isolated test involving training and testing. CD = control diet; HFD = high fat diet; HFHS = high fat-high sugar diet; HSD = high sugar diet).

the bulk of the evidence suggesting a male advantage in spatial navi-


gation (Astur et al., 2016; Moffat et al., 1998; Piber et al., 2018), and a
female advantage in place recognition (Voyer et al., 2007, 2017). Such
differences have been suggested to be due sex hormones (reviewed in
Brake and Lacasse, 2018; Ervin and Choleris, 2019; Hammes and Levin,
2019), such as the differential role of estradiol on hippocampal synaptic
Diet duration Before

plasticity in males and females (Bäumler et al., 2019; Jain et al., 2019;
Wang et al., 2018). Males have typically been used to examine dietary
effects on hippocampal-dependent tasks. When females have been in-
Behaviour

40 days+

10 days+

40 days+

cluded, the experimental designs often lack direct comparison between


8 weeks

8 weeks
17 days

10 days

the sexes (e.g. use of female subjects only, or collapsing across male and
female data points for analysis). Of the 41 articles included in the
present meta-analysis, five articles used female subjects across six ex-
Age* at Diet Onset

periments (Abbott et al., 2016; Jamshed et al., 2014; Kendig et al.,


7+ weeks* (275-

10 months old

2013; Molteni et al., 2004; Pratchayasakul et al., 2015), with deficits


observed in object location memory (Abbott et al., 2016) and in
300 grams)
8 weeks

8 weeks

learning (Kendig et al., 2013) and remembering the platform location


(Jamshed et al., 2014; Molteni et al., 2004) in the Morris water maze
task. Two of these studies collapsed male and female data for analysis
HFHS (labelled H igh-fat, high-dextrose

(Jamshed et al., 2014; Kendig et al., 2013), and another two studies
used only female subjects (Molteni et al., 2004; Pratchayasakul et al.,
2015). Of the included studies, only Abbott et al. (2016) specifically
compared the influence of a high-sugar diet on the performance of male
“Western” Diet by authors)

and female rats in place recognition and object-in-place memory tasks.


They reported that both males and females were impaired in the place
Experimental Diet

recognition task, and males but not females were impaired on object-in-
Ketogenic Diet
Moderate HFD
Extreme HFD

place memory, perhaps indicating that female rats have better memory
for object-in-place configurations than male rats (e.g. Cost et al., 2012).
HFD

4.4. Age as a variable for diet-induced cognitive impairment


ICR mice [Male]
Species & Strain

Sprague Dawley
C57BL/6 J mice

Adolescents may be more susceptible than adults to the effects of


dietary manipulations on hippocampal-dependent learning and
rats [Male]

memory (reviewed in Noble and Kanoski, 2016) and hippocampal


[Male]
[sex]

morphology and function in general (Del Olmo and Ruiz Gayo, 2018).
The transition from adolescence to adulthood occurs at approximately
postnatal day 63 (9 weeks) in rats (Sengupta, 2011, 2013) and postnatal
Hargrave et al (2016)

day 70 (10 weeks) in mice (Dutta and Sengupta, 2016). Using these
Arnold et al (2014)

Park et al (2018)

ages as a cut-off, we extracted 27 experiments from 24 articles that


Author & Expt

manipulated the diet during adolescence (i.e., when aged less than 9
Expt 1

weeks in rats and less than 10 weeks in mice), and nine experiments
Table 8

extracted from nine articles that provided dietary manipulations to


subjects in adulthood. The remaining eight articles combined both

414
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

Fig. 2. Forest plots and meta-analysis results of experiments using a high fat diet. (6-armWM = radial 6-arm water maze; F = female; ExtHF = extreme high fat diet;
HF = high fat diet; KetoDiet = ketogenic diet; M = male; ModHF = moderate high fat diet; MWM = Morris water maze - dependent variable when not time in
target quadrant (Distfromplat = distance from platform); SA = spontaneous alternation; wks = weeks).

adolescents and adults in their control and dietary manipulation con- dietary manipulations on hippocampal-dependent spatial learning and
ditions (Beilharz et al., 2014, 2018; Denver et al., 2018; Hargrave et al., memory. Nevertheless, future meta-analyses may wish to consider age
2016; Kanoski and Davidson, 2010; Kendig et al., 2013; Tran and at the start of the dietary manipulation as a moderator of diet-induced
Westbrook, 2015, 2018). Deficits in hippocampal-dependent spatial hippocampal impairment.
learning and memory was observed in 15 experiments (56%) using
adolescent animals (Abbott et al., 2016; Alzoubi et al., 2013a, 2013b, 4.5. Body weight, metabolic changes, and diet-induced cognitive
2018; Arnold et al., 2014; Jamshed et al., 2014; Liang et al., 2015; impairments
Mirzaei et al., 2018; Molteni et al., 2004; Pathan et al., 2008; Spinelli
et al., 2017; Valladolid-Acebes et al., 2011, 2013, Experiment 1; Xu Our meta-analyses focused on the composition of the diet and the
et al., 2018; Xu and Reichelt, 2018) and in five experiments (56%) task used to assess the effects of the diet on cognition. However, we now
using adults (Che et al., 2018; Moreira et al., 2014; Park et al., 2018; consider the evidence from the included studies regarding the effects of
Thirumangalakudi et al., 2008). Nine experiments failed to detect im- the diet on body weight and metabolic changes (see Table 1). Of the 32
pairments in adolescents (Ayabe et al., 2018; Magnusson et al., 2015; experiments using a HFD: nine experiments did not provide information
Ross et al., 2013; Scichilone et al., 2016; Silva et al., 2005; Takechi regarding body weight, 18 reported a significant difference in body
et al., 2017; Woodie and Blythe, 2018; Wu et al., 2004) and three failed weight between control and experimental animals at the time of be-
to detect impairments in adult (Boitard et al., 2014; Pratchayasakul havioral testing (Alzoubi et al., 2013a, 2013b; Arnold et al., 2014;
et al., 2015; Spencer et al., 2017). Mixed results were reported in three Ayabe et al., 2018; Boitard et al., 2014; Denver et al., 2018; Jurdak
experiments using adolescents (Abbott et al., 2016; Jurdak et al., 2008; et al., 2008; Magnusson et al., 2015; Park et al., 2018; Pathan et al.,
Valladolid-Acebes et al., 2013, Experiment 3) and one experiment using 2008; Spencer et al., 2017; Valladolid-Acebes et al., 2011, 2013; Wei
adults (Gergerlioglu et al., 2016). Thus, the findings from the articles et al., 2018; Woodie and Blythe, 2018; Xu et al., 2018); four reported no
included in the present analysis suggests that the use of adolescent or significant difference in body weight (Gergerlioglu et al., 2016;
adult subjects is not a critical factor in determining the impact of Magnusson et al., 2015; Moreira et al., 2014; Scichilone et al., 2016);

Fig. 3. Forest plots and meta-analysis results of experiments using a high sugar diet. (F = female; HFruc = high fructose diet; HS = high sugar diet; M = male;
MWM = Morris water maze - dependent variable when not time in target quadrant (Distfromplat = distance from platform); wks = weeks).

415
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

Fig. 4. Forest plots and meta-analysis results of experiments using a high fat-high sugar diet. (F = female; HFHFruc = high fat high fructose diet; HFHS = high fat-
high sugar diet; M = male; MWM = Morris water maze; SA = spontaneous alternation; wks = weeks).

Fig. 5. Forest plots and meta-analysis results of experiments using a Morris water maze (MWM). All studies used time in target quadrant as the depdent variable,
unless otherwise stated. (Distfromplat = distance from platform; F = female; HF =high fat diet; HFruc = high fructose diet; HFHFruc = high fat-high fructose diet;
HFHS = high fat-high sugar diet; HS = high sugar diet; M = male; wks = weeks).

Fig. 6. Forest plots and meta-analysis results of experiments using a place recognition task. (F = female; HF = high fat diet; HFHS = high fat-high sufar diet; HS =
high sugar diet; M = male; wks = weeks).

and one reported a significant difference in body weight for rats con- despite intact performance during the probe test for the original plat-
suming the diet from 8 weeks of age, but not for rats consuming the diet form location (Magnusson et al., 2015; Woodie and Blythe, 2018); three
from 4 weeks (Pratchayasakul et al., 2015). Of the experiments re- reported deficits in spontaneous alternation (Arnold et al., 2014, Ex-
porting a significant difference in body weight between control and treme HFD & Moderate HFD; Park et al., 2018); and one reported a
experimental animals: three reported a deficit on the RAM or RAWM deficit in place recognition (Valladolid-Acebes et al., 2013, Experiment
(Alzoubi et al., 2013a, 2013b; Valladolid-Acebes et al., 2011); three 1). The remaining experiments that reported body weight differences
reported a deficit on the MWM (Pathan et al., 2008; Wei et al., 2018; Xu failed to detect an effect of diet manipulation on the MWM (Boitard
et al., 2018); two reported a deficit on the reversal stage of the MWM et al., 2014; Denver et al., 2018; Jurdak et al., 2008; Magnusson et al.,

416
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

Fig. 7. Forest plots and meta-analysis results of experiments using a radial arm or radial arm water maze (WM). (HF = high fat diet; HFHS = high fat-high sugar diet;
M = male; wks = weeks).

Fig. 8. Forest plots and meta-analysis results of experiments using a spontaneous alternation (SA) task. (HF = high fat diet; HFHS = high fat-high sugar diet; M =
male; wks = weeks).

2015; Pratchayasakul et al., 2015; Spencer et al., 2017; Woodie and spontaneous alternation (Hargrave et al., 2016; HFHS 40 days, Keto-
Blythe, 2018), in place recognition (Ayabe et al., 2018; Valladolid- genic 10 and 40 days), or the MWM (Takechi et al., 2017). Taken to-
Acebes et al., 2013, Experiment 3), or in spontaneous alternation gether, the experiments included in the present meta-analysis provide
(Hargrave et al., 2016). Of the studies reporting no significant differ- equivocal evidence regarding the association between dietary-induced
ence in body weight: one reported a deficit on the MWM (Gergerlioglu increases in body weight and cognitive performance; increases in body
et al., 2016) and one reported a deficit in Place recognition (Moreira weight and cognitive deficits do co-occur but deficits have been de-
et al., 2014); the remaining two studies reported no behavioural deficit tected in the absence of increases in body weight and cognition has
on the MWM (Scichilone et al., 2016) or in place recognition been intact in spite of increases in body weight.
(Magnusson et al., 2015). The studies included in the meta-analysis differed with respect to
Of the nine experiments using a HSD, only one experiment (Jurdak their assessment of diet-induced alterations to metabolic status. A
et al., 2008) reported a significant difference in body weight between number of studies demonstrating diet-induced cognitive deficits report
control and experimental animals at the time of behavioural testing and significant differences between control and experimental animals in
a dietary-induced deficit in MWM. The remaining eight experiments metabolic parameters including blood glucose level (Arnold et al.,
reported no significant difference in body weight (Abbott et al., 2016; 2014; Jurdak et al., 2008; Pathan et al., 2008; Xu et al., 2018), response
Beilharz et al., 2014; Gergerlioglu et al., 2016; Magnusson et al., 2015). on a glucose tolerance test (Arnold et al., 2014; Xu et al., 2018) and
Of these reporting no difference in body weight: four experiments re- insulin tolerance test (Xu et al., 2018), plasma insulin level (Beilharz
ported a deficit in place recognition (Abbott et al., 2016; Beilharz et al., et al., 2014, Experiment 1; Pathan et al., 2008; Spinelli et al., 2017;
2014; Xu and Reichelt, 2018); one reported a deficit in the MWM Valladolid-Acebes et al., 2011), plasma leptin level (Beilharz et al.,
(Kendig et al., 2013). The remaining experiments reported no deficit in 2014, Experiment 2; Valladolid-Acebes et al., 2013, Experiment 1),
place recognition (Magnusson et al., 2015) or the MWM (Gergerlioglu triglycerides (Beilharz et al., 2014, Experiment 2; Jamshed et al., 2014;
et al., 2016; Magnusson et al., 2015). Jurdak et al., 2008; Moreira et al., 2014; Pathan et al., 2008; Xu et al.,
Of the 11 studies using a HFSD, four reported a significant differ- 2018), and cholesterol levels (Jamshed et al., 2014; Moreira et al.,
ence in body weight between control and experimental animals at the 2014; Pathan et al., 2008; Xu et al., 2018). However, a number of the
time of behavioral testing (Beilharz et al., 2018; Kanoski and Davidson, included studies also failed to find statistically significant differences in
2010; Ross et al., 2013; Takechi et al., 2017); one reported no sig- these metabolic parameters (Beilharz et al., 2014, 2018; Gergerlioglu
nificant difference in body weight (Woodie and Blythe, 2018); one re- et al., 2016; Jurdak et al., 2008; Kendig et al., 2013; Moreira et al.,
ported a significant difference in total body fat but not body weight 2014; Valladolid-Acebes et al., 2011, 2013). Furthermore, dietary-in-
(Hargrave et al., 2016); and three reported a significant difference in duced alterations in blood glucose (Boitard et al., 2014; Kendig et al.,
body weight at later, but not early, behavioural tests (Beilharz et al., 2013; Scichilone et al., 2016; Takechi et al., 2017; Valladolid-Acebes
2014, Experiments 1 and 2; Tran and Westbrook, 2015). Of the studies et al., 2013, Experiment 3), plasma insulin and HOMA (Pratchayasakul
reporting a significant difference in body fat or body weight at one or et al., 2015; Takechi et al., 2017), and plasma leptin (Boitard et al.,
more behavioural test point: four reported a deficit in place recognition 2014; Valladolid-Acebes et al., 2013, Experiment 3) have been reported
(Beilharz et al., 2014, Expeirments 1 and 2; Beilharz et al., 2018; Tran but in the absence of deficits in hippocampal-dependent learning and
and Westbrook, 2015): one reported a deficit in spontaneous alterna- memory. Thus, again, the experiments included in the meta-analysis
tion (Hargrave et al., 2016; HFHS 10 days); and one reported a deficit in provide equivocal evidence regarding the association between meta-
the RAM (Kanoski and Davidson, 2010). The remaining experiments bolic changes and cognitive performance.
reported no deficits in place recognition (Ross et al., 2013),

417
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

4.6. Mechanisms of diet-induced impairments type of meta-analysis conducted, the weighted effect-sizes suggest that
diet-induced impairments are most likely to be detected in rodents fed a
There are various mechanisms by which high fat, high sugar, or high high fat-high sugar diet and tested in the radial arm maze.
fat-high sugar diets can impair hippocampal-dependent forms of cog-
nition (reviewed in Beilharz et al., 2015; Freeman et al., 2014). One Acknowledgement
such mechanism is neuroinflammation (Guillemot-Legris, and Muccioli,
2017). Several studies have reported that diet-induced impairments are Kirsten Abbott was supported by an Australian Government
accompanied by increased expression of pro-inflammatory cytokines, Research Training Program Scholarship.
such as interleukin-1β (IL-1β; Beilharz et al., 2014, 2018; Che et al.,
2018; Mirzaei et al., 2018; Thirumangalakudi et al., 2008), tumor ne- Appendix A
crosis factor-α (TNF-α) mRNA (Beilharz et al., 2014; Thirumangalakudi
et al., 2008) and protein (Che et al., 2018), as well as glial fibrillary Search Terms and Strategy
acidic protein (GFAP), a marker of glial cell activation, (Che et al., A search was conducted on PubMed for the following terms:
2018; Thirumangalakudi et al., 2008). However, inconsistent results High Fat Diet:
have been reported in rodents who failed to exhibit diet-induced im- (((((((((high fat[Title/Abstract]) AND memory[Title/Abstract]) OR
pairments: some studies failed to detect evidence for pro-inflammatory fat[Title/Abstract]) AND diet[Title/Abstract]) AND memory[Title/
cytokine expression (Boitard et al., 2014; Takechi et al., 2017) and Abstract]) OR high fat[Title/Abstract]) AND cognit*[Title/Abstract])
GFAP (Denver et al., 2018; Silva et al., 2005), whereas others found OR fat[Title/Abstract]) AND diet[Title/Abstract]) AND cognit*[Title/
increased expression of IL-1β and TNF- α protein (Ayabe et al., 2018; Abstract]
Spencer et al., 2017). Thus, the presence of neuroinflammation is not a High Sugar Diet:
necessarily condition for observing short-term dietary impairments in ((((((((high sugar[Title/Abstract]) AND memory[Title/Abstract])
hippocampal-dependent spatial learning and memory. OR sugar[Title/Abstract]) AND diet[Title/Abstract]) AND memory
A second mechanism is reductions in neuroplasticity (Morin et al., [Title/Abstract]) OR high sugar[Title/Abstract]) AND cognit*[Title/
2017; Murphy et al., 2014). Several studies have reported that diet- Abstract]) OR sugar[Title/Abstract]) AND cognit*[Title/Abstract]
induced impairments are associated with reduced long-term potentia- High Fat Sugar Diet:
tion in hippocampal pyramidal neurons (Spinelli et al., 2017) and (((((((Western diet) AND memory) OR cafeteria diet[Title/
changes in factors related to neuroplasticity, such as decreases in sy- Abstract]) AND memory[Title/Abstract]) OR Western diet[Title/
napsin-1 mRNA and protein (Molteni et al., 2004), as well as in Brain Abstract]) AND cognit*[Title/Abstract]) OR cafeteria diet[Title/
Derived Neurotrophic Factor (BDNF) mRNA and protein (Che et al., Abstract]) AND cognit*[Title/Abstract]
2018; Molteni et al., 2004). Consistent with a role for changes in neu- A search on Proquest was conducted using the search terms:
roplasticity, some studies failed to detect any changes in BDNF mRNA High Fat Diet:
(Wu et al., 2003) or BDNF protein expression (Woodie and Blythe, ab(high fat) AND ab(memory) OR ab(fat) AND ab(diet) AND ab
2018; Wu et al., 2003) in rodents who failed to exhibit diet-induced (memory) OR ab(high fat) AND ab(cognit*) OR ab(fat) AND ab(diet)
impairments in spatial learning and memory. However, other studies AND ab(cognit*)
have reported diet-induced impairments, but in the absence of changes High Sugar Diet:
to BDNF mRNA (Alzoubi et al., 2013a, 2013b) or BDNF protein ex- ab(high sugar) AND ab(memory) OR ab(sugar) AND ab(diet) AND
pression (Beilharz et al., 2014). Conversely, markers of altered neuro- ab(memory) OR ab(high sugar) AND ab(cognit*) OR ab(sugar) AND ab
plasticity, such as increased synaptophysin protein (Denver et al., (diet) AND ab(cognit*)
2018), decreased dendritic spine density, as well as both increases High Fat Sugar Diet:
(Ayabe et al., 2018) and decreases (Scichilone et al., 2016) in BDNF ab(western diet) AND ab(memory) OR ab(cafeteria diet) AND ab
protein expression, have been found in rodents who failed to exhibit (memory) OR ab(Western diet) AND ab(cognit*) OR ab(cafeteria diet)
diet-induced impairments in spatial learning and memory. Taken to- AND ab(cognit*)
gether, changes in neuroplasticity may play a role in mediating the
cognitive impairments induced by relatively short-term dietary ex- References
posures but the nature of this role remains to be determined.
Finally, additional mechanisms include altered insulin signaling Abbott, K.N., Morris, M.J., Westbrook, R.F., Reichelt, A.C., 2016. Sex-specific effects of
within the hippocampus (Arnold et al., 2014; Liang et al., 2015; Spinelli daily exposure to sucrose on spatial memory performance in male and female rats,
and implications for estrous cycle stage. Physiol. Behav. 162, 52–60. https://doi.org/
et al., 2017; Xu et al., 2018), disruptions to the blood-brain-barrier (Hsu 10.1016/j.physbeh.2016.01.036.
and Kanoski, 2014), and dysbiosis in the gut microbiome arising from Alzoubi, K.H., Khabour, O.F., Salah, H.A., Abu Rashid, B.E., 2013a. The combined effect
reduced bacterial diversity (Beilharz et al., 2018; Magnusson et al., of sleep deprivation and Western diet on spatial learning and memory: role of BDNF
and oxidative stress. J. Mol. Neurosci. 50 (1), 124–133. https://doi.org/10.1007/
2015; Noble et al., 2017). However, these proposals are relatively re- s12031-012-9881-7.
cent and they await further investigation (Denver et al., 2018; Alzoubi, K.H., Khabour, O.F., Salah, H.A., Hasan, Z., 2013b. Vitamin E prevents high- fat
Pratchayasakul et al., 2015). high-carbohydrates diet-induced memory impairment: the role of oxidative stress.
Physiol. Behav. 119, 72–78. https://doi.org/10.1016/j.physbeh.2013.06.011.
Alzoubi, K.H., Mayyas, F.A., Mahafzah, R., Khabour, O.F., 2018. Melatonin prevents
5. Conclusion memory impairment induced by high-fat diet: role of oxidative stress. Behav. Brain
Res. 336, 93–98. https://doi.org/10.1016/j.bbr.2017.08.047.
Arnold, S.E., Lucki, I., Brookshire, B.R., Carlson, G.C., Browne, C.A., Kazi, H., et al., 2014.
In sum, the meta-analysis showed that, despite the inconsistent
High fat diet produces brain insulin resistance, synaptodendritic abnormalities and
findings in the literature, diets high in saturated fat or refined sugar can altered behavior in mice. Neurobiol. Dis. 67, 79–87. https://doi.org/10.1016/j.nbd.
each produce reliable impairments in hippocampal-dependent forms of 2014.03.011.
cognition. It also showed that these impairments can be detected in a Astur, R.S., Purton, A.J., Zaniewski, M.J., Cimadevilla, J., Markus, E.J., 2016. Human sex
differences in solving a virtual navigation problem. Behav. Brain Res. 308, 236–243.
range of hippocampal-dependent tasks. These results reveal that the https://doi.org/10.1016/j.bbr.2016.04.037.
type of dietary manipulation is unlikely to be the primary factor for the Attuquayefio, T., Stevenson, R.J., Oaten, M.J., Francis, H.M., 2017. A four-day Western-
inconsistent findings in the literature. Instead, we propose that the in- style dietary intervention causes reductions in hippocampal-dependent learning and
memory and interoceptive sensitivity. PLoS One 12 (2). https://doi.org/10.1371/
consistencies are more likely to be a product of the procedural varia- journal.pone.0172645.
tions used to assess the impact of short-term diet exposures on cogni- Ayabe, T., Ohya, R., Kondo, K., Ano, Y., 2018. Iso-α-acids, bitter components of beer,
tion. Finally, while no direct statistical comparisons were made in this prevent obesity-induced cognitive decline. Sci. Rep. 8 (1), 4760. https://doi.org/10.

418
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

1038/s41598-018-23213-9. intake is associated with reduced hippocampal-dependent memory and sensitivity to


Barker, G.R., Warburton, E.C., 2011. When is the hippocampus involved in recognition interoceptive signals. Behav. Neurosci. 125 (6), 943. https://doi.org/10.1037/
memory? J. Neurosci. 31 (29), 10721–10731. https://doi.org/10.1523/JNEUROSCI. a0025998.
6413-10.2011. Freeman, L.R., Haley-Zitlin, V., Rosenberger, D.S., Granholm, A.C., 2014. Damaging ef-
Bäumler, E., Strickland, L., Privitera, L., 2019. Molecular underpinnings of estradiol- fects of a high-fat diet to the brain and cognition: a review of proposed mechanisms.
mediated sexual dimorphism of synaptic plasticity in the hippocampus of rodents. J. Nutr. Neurosci. 17 (6), 241–251. https://doi.org/10.1179/1476830513Y.
Neurosci. 39 (12), 2160–2162. 0000000092.
Beilharz, J.E., Maniam, J., Morris, M.J., 2014. Short exposure to a diet rich in both fat and Genzel, L., Schut, E., Schroeder, T., Eichler, R., Bayraktar, G., Cornelisse, N., et al., 2017.
sugar or sugar alone impairs place, but not object recognition memory in rats. Brain The object space task for mice and rats. bioRxiv 198382. https://doi.org/10.1101/
Behav. Immun. 37, 134–141. https://doi.org/10.1016/j.bbi.2013.11.016. 198382.
Beilharz, J., Maniam, J., Morris, M., 2015. Diet-induced cognitive deficits: the role of fat Gergerlioglu, H.S., Oz, M., Demir, E.A., Nurullahoglu-Atalik, K.E., Yerlikaya, F.H., 2016.
and sugar, potential mechanisms and nutritional interventions. Nutrients 7 (8), Environmental enrichment reverses cognitive impairments provoked by Western diet
6719–6738. https://doi.org/10.3390/nu7085307. in rats: role of corticosteroid receptors. Life Sci. 148, 279–285. https://doi.org/10.
Beilharz, J.E., Kaakoush, N.O., Maniam, J., Morris, M.J., 2018. Cafeteria diet and pro- 1016/j.lfs.2016.02.011.
biotic therapy: cross talk among memory, neuroplasticity, serotonin receptors and gut Gibson, E.L., Barr, S., Jeanes, Y.M., 2013. Habitual fat intake predicts memory function in
microbiota in the rat. Mol. Psychiatry 23 (2), 351. https://doi.org/10.1038/mp. younger women. Front. Hum. Neurosci. 7, 838. https://doi.org/10.3389/fnhum.
2017.38. 2013.00838.
Boitard, C., Cavaroc, A., Sauvant, J., Aubert, A., Castanon, N., Laye, S., Ferreira, G., 2014. Guillemot-Legris, O., Muccioli, G.G., 2017. Obesity-induced neuroinflammation: beyond
Impairment of hippocampal-dependent memory induced by juvenile high-fat diet the hypothalamus. Trends Neurosci. 40 (4), 237–253. https://doi.org/10.1016/j.tins.
intake is associated with enhanced hippocampal inflammation in rats. Brain Behav. 2017.02.005.
Immun. 40, 9–17. https://doi.org/10.1016/j.bbi.2014.03.005. Gustaw-Rothenberg, K., 2009. Dietary patterns associated with Alzheimer’s disease: po-
Brake, W.G., Lacasse, J.M., 2018. Sex differences in spatial navigation: the role of gonadal pulation based study. Int. J. Environ. Res. Public Health 6 (4), 1335–1340. https://
hormones. Curr. Opin. Behav. Sci. 23, 176–182. https://doi.org/10.1016/j.cobeha. doi.org/10.3390/ijerph6041335.
2018.08.002. Hammes, S.R., Levin, E.R., 2019. Impact of estrogens in males and androgens in females.
Buettner, R., Schölmerich, J., Bollheimer, L.C., 2007. High-fat diets: modeling the me- J. Clin. Invest. 129 (5), 1818–1826. https://doi.org/10.1172/JCI125755.
tabolic disorders of human obesity in rodents. Obesity 15 (4), 798–808. https://doi. Hargrave, S.L., Davidson, T.L., Zheng, W., Kinzig, K.P., 2016. Western diets induce blood-
org/10.1038/oby.2007.608. brain barrier leakage and alter spatial strategies in rats. Behav. Neurosci. 130 (1),
Carrera-Bastos, P., Fontes-Villalba, M., O’Keefe, J.H., Lindeberg, S., Cordain, L., 2011. The 123–135. https://doi.org/10.1037/bne0000110.
western diet and lifestyle and diseases of civilization. Res. Rep. Clin. Cardiol. 2, Hedges, L.V., 1981. Distribution theory for Glass’s estimator of effect size and related
15–35. https://doi.org/10.2147/RRCC.S16919. estimators. J. Educ. Stat. 6 (2), 107–128.
Chang, S., Graham, B., Yakubu, F., Lin, D., Peters, J.C., Hill, J.O., 1990. Metabolic dif- Hok, V., Poucet, B., Duvelle, É., Save, É., Sargolini, F., 2016. Spatial cognition in mice and
ferences between obesity-prone and obesity-resistant rats. Am. J. Physiol.- rats: similarities and differences in brain and behavior. Wiley Interdiscip. Rev. Cogn.
Regul.Integr. Comparative Physiol. 259 (6), R1103–R1110. https://doi.org/10.1152/ Sci. 7 (6), 406–421. https://doi.org/10.1002/wcs.1411.
ajpregu.1990.259.6.R1103. Hsu, T.M., Kanoski, S.E., 2014. Blood-brain barrier disruption: mechanistic links between
Che, H., Zhou, M., Zhang, T., Zhang, L., Ding, L., Yanagita, T., et al., 2018. Comparative Western diet consumption and dementia. Front. Aging Neurosci. 6, 88. https://doi.
study of the effects of phosphatidylcholine rich in DHA and EPA on Alzheimer’s org/10.3389/fnagi.2014.00088.
disease and the possible mechanisms in CHO-APP/PS1 cells and SAMP8 mice. Food Jain, A., Huang, G.Z., Woolley, C.S., 2019. Latent sex differences in molecular signaling
Funct. 9 (1), 643–654. https://doi.org/10.1039/C7FO01342F. that underlies excitatory synaptic potentiation in the hippocampus. J. Neurosci. 39
Compton, D.M., Griffith, H.R., McDaniel, W.F., Foster, R.A., Davis, B.K., 1997. The flex- (9), 1552–1565. https://doi.org/10.1523/JNEUROSCI.1897-18.2018.
ible use of multiple cue relationships in spatial navigation: a comparison of water Jamshed, H., Arslan, J., Gilani, A.-u.-H., 2014. Cholesterol-cholate-butterfat diet offers
maze performance following hippocampal, medial septal, prefrontal cortex, or pos- multi-organ dysfunction in rats. Lipids Health Dis. 13. https://doi.org/10.1186/
terior parietal cortex lesions. Neurobiol. Learn. Mem. 68 (2), 117–132. https://doi. 1476-511X-13-194.
org/10.1006/nlme.1997.3793. Ji, D., Mou, X., Cheng, J., Yu, Y.S., Kee, S.E., 2018. Comparing mouse and rat hippo-
Cordain, L., Eaton, S.B., Sebastian, A., Mann, N., Lindeberg, S., Watkins, B.A., et al., 2005. campal place cell activities and firing sequences in the same environments. Front.
Origins and evolution of the Western diet: health implications for the 21st century. Cell. Neurosci. 12, 332. https://doi.org/10.3389/fncel.2018.00332.
Am. J. Clin. Nutr. 81 (2), 341–354. https://doi.org/10.1093/ajcn.81.2.341. Jurdak, N., Lichtenstein, A.H., Kanarek, R.B., 2008. Diet-induced obesity and spatial
Cordner, Z.A., Tamashiro, K.L., 2015. Effects of high-fat diet exposure on learning & cognition in young male rats. Nutr. Neurosci. 11 (2), 48–54. https://doi.org/10.
memory. Physiol. Behav. 152, 363–371. https://doi.org/10.1016/j.physbeh.2015.06. 1179/147683008X301333.
008. Kang, S.S., Jeraldo, P.R., Kurti, A., Miller, M.E.B., Cook, M.D., Whitlock, K., et al., 2014.
Cost, K.T., Williams-Yee, Z.N., Fustok, J.N., Dohanich, G.P., 2012. Sex differences in Diet and exercise orthogonally alter the gut microbiome and reveal independent
object-in-place memory of adult rats. Behav. Neurosci. 126 (3), 457. associations with anxiety and cognition. Mol. Neurodegener. 9 (1), 36. https://doi.
Cressant, A., Besson, M., Suarez, S., Cormier, A., Granon, S., 2007. Spatial learning in org/10.1186/1750-1326-9-36.
Long-Evans Hooded rats and C57BL/6J mice: different strategies for different per- Kanoski, S.E., Davidson, T.L., 2010. Different patterns of memory impairments accom-
formance. Behav. Brain Res. 177 (1), 22–29. https://doi.org/10.1016/j.bbr.2006.11. pany short- and longer-term maintenance on a high-energy diet. J. Exp. Psychol.
010. Anim. Behav. Process. 36 (2), 313–319. https://doi.org/10.1037/a0017228.
Davidson, T.L., Hargrave, S.L., Swithers, S.E., Sample, C.H., Fu, X., Kinzig, K.P., Zheng, Kanoski, S.E., Davidson, T.L., 2011. Western diet consumption and cognitive impairment:
W., 2013. Inter-relationships among diet, obesity and hippocampal-dependent cog- links to hippocampal dysfunction and obesity. Physiol. Behav. 103 (1), 59–68.
nitive function. Neuroscience 253, 110–122. https://doi.org/10.1016/j. https://doi.org/10.1016/j.physbeh.2010.12.003.
neuroscience.2013.08.044. Kanoski, S.E., Meisel, R.L., Mullins, A.J., Davidson, T.L., 2007. The effects of energy-rich
Del Olmo, N., Ruiz Gayo, M., 2018. Influence of high-fat diets consumed during the ju- diets on discrimination reversal learning and on BDNF in the hippocampus and
venile period on hippocampal morphology and function. Front. Cell. Neurosci. 12, prefrontal cortex of the rat. Behav. Brain Res. 182 (1), 57–66. https://doi.org/10.
439. https://doi.org/10.3389/fncel.2018.00439. 1016/j.bbr.2007.05.004.
Denver, P., Gault, V.A., McClean, P.L., 2018. Sustained high-fat diet modulates in- Kanoski, S.E., Zhang, Y., Zheng, W., Davidson, T.L., 2010. The effects of a high- energy
flammation, insulin signalling and cognition in mice and a modified xenin peptide diet on hippocampal function and blood-brain barrier integrity in the rat. J.
ameliorates neuropathology in a chronic high-fat model. Diabetes Obes. Metab. 20 Alzheimers Dis. 21 (1), 207–219. https://doi.org/10.3233/jad-2010-091414.
(5), 1166–1175. https://doi.org/10.1111/dom.13210. Kearney, J., 2010. Food consumption trends and drivers. Philos. Trans. Biol. Sci. 365
Dutta, S., Sengupta, P., 2016. Men and mice: relating their ages. Life Sci. 152, 244–248. (1554), 2793–2807. https://doi.org/10.1098/rstb.2010.0149.
https://doi.org/10.1016/j.lfs.2015.10.025. Keeley, R.J., Bye, C., Trow, J., McDonald, R.J., 2015. Strain and sex differences in brain
Ennaceur, A., Michalikova, S., Bradford, A., Ahmed, S., 2005. Detailed analysis of the and behaviour of adult rats: Learning and memory, anxiety and volumetric estimates.
behavior of Lister and Wistar rats in anxiety, object recognition and object location Behav. Brain Res. 288, 118–131. https://doi.org/10.1016/j.bbr.2014.10.039.
tasks. Behav. Brain Res. 159 (2), 247–266. https://doi.org/10.1016/j.bbr.2004.11. Kendig, M.D., Boakes, R.A., Rooney, K.B., Corbit, L.H., 2013. Chronic restricted access to
006. 10% sucrose solution in adolescent and young adult rats impairs spatial memory and
Ervin, K.S., Choleris, E., 2019. Involvement of the sex hormones in learning and memory. alters sensitivity to outcome devaluation. Physiol. Behav. 120, 164–172. https://doi.
Oxford Handbook Evolutionary Psychol. Behav. Endocrinol. 67. https://doi.org/10. org/10.1016/j.physbeh.2013.08.012.
1093/oxfordhb/9780190649739.013.4. Khatibzadeh, S., Kashaf, M.S., Micha, R., Fahimi, S., Shi, P., Elmadfa, I., et al., 2016. A
Eskelinen, M.H., Ngandu, T., Helkala, E.L., Tuomilehto, J., Nissinen, A., Soininen, H., global database of food and nutrient consumption. Bull. World Health Organ. 94 (12),
Kivipelto, M., 2008. Fat intake at midlife and cognitive impairment later in life: a 931. https://doi.org/10.2471/BLT.15.156323.
population-based CAIDE study. Int. J. Geriatric Psychiatry 23 (7), 741–747. https:// Koss, W.A., Frick, K.M., 2017. Sex differences in hippocampal function. J. Neurosci. Res.
doi.org/10.1002/gps.1969. 95 (1-2), 539–562. https://doi.org/10.1002/jnr.23864.
Fan, S.J., Sun, A.B., Liu, L., 2018. Epigenetic modulation during hippocampal develop- Lalonde, R., 2002. The neurobiological basis of spontaneous alternation. Neurosci.
ment. Biomed. Rep. 9 (6), 463–473. https://doi.org/10.3892/br.2018.1160. Biobehav. Rev. 26 (1), 91–104. https://doi.org/10.1016/S0149-7634(01)00041-0.
Floresco, S.B., Seamans, J.K., Phillips, A.G., 1997. Selective roles for hippocampal, pre- Liang, L., Chen, J., Zhan, L., Lu, X., Sun, X., Sui, H., et al., 2015. Endoplasmic reticulum
frontal cortical, and ventral striatal circuits in radial-arm maze tasks with or without stress impairs insulin receptor signaling in the brains of obese rats. PLoS One 10 (5),
a delay. J. Neurosci. 17 (5), 1880–1890. https://doi.org/10.1523/JNEUROSCI.17- e0126384. https://doi.org/10.1371/journal.pone.0126384.
05-01880.1997. Magnusson, K.R., Hauck, L., Jeffrey, B.M., Elias, V., Humphrey, A., Nath, R., et al., 2015.
Francis, H.M., Stevenson, R.J., 2011. Higher reported saturated fat and refined sugar Relationships between diet-related changes in the gut microbiome and cognitive

419
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

flexibility. Neuroscience 300, 128–140. https://doi.org/10.1016/j.neuroscience. Molecular Basis Disease 1863 (5), 1037–1045. https://doi.org/10.1016/j.bbadis.
2015.05.016. 2016.04.017.
Malki, K., Mineur, Y.S., Tosto, M.G., Campbell, J., Karia, P., Jumabhoy, I., et al., 2015. Radic, T., Frieß, L., Vijikumar, A., Jungenitz, T., Deller, T., Schwarzacher, S.W., 2017.
Pervasive and opposing effects of Unpredictable Chronic Mild Stress (UCMS) on Differential postnatal expression of neuronal maturation markers in the dentate gyrus
hippocampal gene expression in BALB/cJ and C57BL/6J mouse strains. BMC of mice and rats. Front. Neuroanat. 11, 104. https://doi.org/10.3389/fnana.2017.
Genomics 16 (1), 262. https://doi.org/10.1186/s12864-015-1431-6. 00104.
Manahan-Vaughan, D., 2019. Special considerations when using mice for in vivo elec- Ritchie, H., Roser, M., 2017. Diet Compositions. Retrieved from https//ourworldinda-
trophysiology and long-term studies of hippocampal synaptic plasticity during be- ta.org/diet-compositions. .
havior. Handbook Behav. Neurosci. 28, 63–84. https://doi.org/10.1016/B978-0-12- Roberts, R.O., Roberts, L.A., Geda, Y.E., Cha, R.H., Pankratz, V.S., O’Connor, H.M., et al.,
812028-6.00003-3. Elsevier. 2012. Relative intake of macronutrients impacts risk of mild cognitive impairment or
Mente, A., de Koning, L., Shannon, H.S., Anand, S.S., 2009. A systematic review of the dementia. J. Alzheimer Dis. 32 (2), 329–339. https://doi.org/10.3233/JAD-2012-
evidence supporting a causal link between dietary factors and coronary heart disease. 120862.
Arch. Intern. Med. 169 (7), 659–669. https://doi.org/10.1001/archinternmed. Ross, A.P., Darling, J.N., Parent, M.B., 2013. High energy diets prevent the enhancing
2009.38. effects of emotional arousal on memory. Behav. Neurosci. 127 (5), 771. https://doi.
Merritt, J.R., Rhodes, J.S., 2015. Mouse genetic differences in voluntary wheel running, org/10.1037/a0033732.
adult hippocampal neurogenesis and learning on the multi-strain-adapted plus water Scichilone, J.M., Yarraguntla, K., Charalambides, A., Harney, J.P., Butler, D., 2016.
maze. Behav. Brain Res. 280, 62–71. https://doi.org/10.1016/j.bbr.2014.11.030. Environmental Enrichment Mitigates Detrimental Cognitive Effects of Ketogenic Diet
Mirzaei, F., Khazaei, M., Komaki, A., Amiri, I., Jalili, C., 2018. Virgin coconut oil (VCO) in Weanling Rats. J. Mol. Neurosci. 60 (1), 1–9. https://doi.org/10.1007/s12031-
by normalizing NLRP3 inflammasome showed potential neuroprotective effects in 016-0753-4.
Amyloid-β induced toxicity and high-fat diet fed rat. Food Chem. Toxicol. 118, Semple, B.D., Blomgren, K., Gimlin, K., Ferriero, D.M., Noble-Haeusslein, L.J., 2013.
68–83. https://doi.org/10.1016/j.fct.2018.04.064. Brain development in rodents and humans: identifying benchmarks of maturation
Moffat, S.D., Hampson, E., Hatzipantelis, M., 1998. Navigation in a "virtual" maze: sex and vulnerability to injury across species. Prog. Neurobiol. 106, 1–16. https://doi.
differences and correlation with psychometric measures of spatial ability in humans. org/10.1016/j.pneurobio.2013.04.001.
Evol. Hum. Behav. 19 (2), 73–87. https://doi.org/10.1016/S1090-5138(97)00104-9. Sengupta, P., 2011. A scientific review of age determination for a laboratory rat: how old
Molteni, R., Barnard, R.J., Ying, Z., Roberts, C.K., Gomez-Pinilla, F., 2002. A high-fat, is it in comparison with human age. Biomed Int 2 (2), 81–89.
refined sugar diet reduces hippocampal brain-derived neurotrophic factor, neuronal Sengupta, P., 2013. The laboratory rat: relating its age with human’s. Int. J. Prev. Med. 4
plasticity, and learning. Neuroscience 112 (4), 803–814. https://doi.org/10.1016/ (6), 624.
S0306-4522(02)00123-9. Silva, M.C., Rocha, J., Pires, C.S., Ribeiro, L.C., Brolese, G., Leite, M.C., et al., 2005.
Molteni, R., Wu, A., Vaynman, S., Ying, Z., Barnard, R., Gomez-Pinilla, F., 2004. Exercise Transitory gliosis in the CA3 hippocampal region in rats fed on a ketogenic diet. Nutr.
reverses the harmful effects of consumption of a high-fat diet on synaptic and be- Neurosci. 8 (4), 259–264. https://doi.org/10.1080/10284150500475032.
havioral plasticity associated to the action of brain-derived neurotrophic factor. Sobesky, J.L., Barrientos, R.M., Henning, S., Thompson, B.M., Weber, M.D., Watkins, L.R.,
Neuroscience 123 (2), 429–440. https://doi.org/10.1016/j.neuroscience.2003.09. Maier, S.F., 2014. High-fat diet consumption disrupts memory and primes elevations
020. in hippocampal IL-1β, an effect that can be prevented with dietary reversal or IL-1
Moreira, E.L., de Oliveira, J., Engel, D.F., Walz, R., de Bem, A.F., Farina, M., Prediger, receptor antagonism. Brain Behav. Immun. 42, 22–32. https://doi.org/10.1016/j.bbi.
R.D., 2014. Hypercholesterolemia induces short-term spatial memory impairments in 2014.06.017.
mice: up-regulation of acetylcholinesterase activity as an early and causal event? J. Spencer, S.J., D’Angelo, H., Soch, A., Watkins, L.R., Maier, S.F., Barrientos, R.M., 2017.
Neural Transm. Vienna (Vienna) 121 (4), 415–426. https://doi.org/10.1007/s00702- High-fat diet and aging interact to produce neuroinflammation and impair hippo-
013-1107-9. campal-and amygdalar-dependent memory. Neurobiol. Aging 58, 88–101. https://
Morin, J.P., Rodríguez-Durán, L.F., Guzmán-Ramos, K., Perez-Cruz, C., Ferreira, G., Diaz- doi.org/10.1016/j.neurobiolaging.2017.06.014.
Cintra, S., Pacheco-López, G., 2017. Palatable hyper-caloric foods impact on neuronal Spinelli, M., Fusco, S., Mainardi, M., Scala, F., Natale, F., Lapenta, R., et al., 2017. Brain
plasticity. Front. Behav. Neurosci. 11, 19. https://doi.org/10.3389/fnbeh.2017. insulin resistance impairs hippocampal synaptic plasticity and memory by increasing
00019. GluA1 palmitoylation through FoxO3a. Nat. Commun. 8 (1), 2009. https://doi.org/
Morris, M.C., Evans, D.A., Bienias, J.L., Tangney, C.C., Wilson, R.S., 2004. Dietary fat 10.1038/s41467-017-02221-9.
intake and 6-year cognitive change in an older biracial community population. Stevens, R., Cowey, A., 1973. Effects of dorsal and ventral hippocampal lesions on
Neurology 62 (9), 1573–1579. https://doi.org/10.1212/01.WNL.0000123250. spontaneous alternation, learned alternation and probability learning in rats. Brain
82849.B6. Res. 52, 203–224. https://doi.org/10.1016/0006-8993(73)90659-8.
Murphy, T., Dias, G.P., Thuret, S., 2014. Effects of diet on brain plasticity in animal and Stöckli, J., Fisher-Wellman, K.H., Chaudhuri, R., Zeng, X.Y., Fazakerley, D.J., Meoli, C.C.,
human studies: mind the gap. Neural Plast. 2014. https://doi.org/10.1155/2014/ et al., 2017. Metabolomic analysis of insulin resistance across different mouse strains
563160. and diets. J. Biol. Chem. 292 (47), 19135–19145. https://doi.org/10.1074/jbc.M117.
Murray, A.J., Knight, N.S., Cochlin, L.E., McAleese, S., Deacon, R.M., Rawlins, J.N., 818351.
Clarke, K., 2009. Deterioration of physical performance and cognitive function in rats Stranahan, A.M., 2011. Similarities and differences in spatial learning and object re-
with short-term high-fat feeding. FASEB J. 23 (12), 4353–4360. https://doi.org/10. cognition between young male C57Bl/6J mice and Sprague-Dawley rats. Behav.
1096/fj.09-139691. Neurosci. 125 (5), 791. https://doi.org/10.1037/a0025133.
Naqvi, A.Z., Harty, B., Mukamal, K.J., Stoddard, A.M., Vitolins, M., Dunn, J.E., 2011. Takechi, R., Lam, V., Brook, E., Giles, C., Fimognari, N., Mooranian, A., et al., 2017.
Monounsaturated, trans, and saturated fatty acids and cognitive decline in women. J. Blood-brain barrier dysfunction precedes cognitive decline and neurodegeneration in
Am. Geriatr. Soc. 59 (5), 837–843. https://doi.org/10.1111/j.1532-5415.2011. diabetic insulin resistant mouse model: an implication for causal link. Front. Aging
03402.x. Neurosci. 9, 399. https://doi.org/10.3389/fnagi.2017.00399.
Noble, E.E., Kanoski, S.E., 2016. Early life exposure to obesogenic diets and learning and Thirumangalakudi, L., Prakasam, A., Zhang, R., Bimonte-Nelson, H., Sambamurti, K.,
memory dysfunction. Curr. Opin. Behav. Sci. 9, 7–14. https://doi.org/10.1016/j. Kindy, M.S., Bhat, N.R., 2008. High cholesterol-induced neuroinflammation and
cobeha.2015.11.014. amyloid precursor protein processing correlate with loss of working memory in mice.
Noble, E.E., Hsu, T.M., Kanoski, S.E., 2017. Gut to brain dysbiosis: mechanisms linking J. Neurochem. 106 (1), 475–485. https://doi.org/10.1111/j.1471-4159.2008.
western diet consumption, the microbiome, and cognitive impairment. Front. Behav. 05415.x.
Neurosci. 11, 9. https://doi.org/10.3389/fnbeh.2017.00009. Tran, D.M., Westbrook, R.F., 2015. Rats fed a diet rich in fats and sugars are impaired in
Okereke, O.I., Rosner, B.A., Kim, D.H., Kang, J.H., Cook, N.R., Manson, J.E., et al., 2012. the use of spatial geometry. Psychol. Sci. 26 (12), 1947–1957. https://doi.org/10.
Dietary fat types and 4-year cognitive change in community-dwelling older women. 1177/0956797615608240.
Ann. Neurol. 72 (1), 124–134. https://doi.org/10.1002/ana.23593. Tran, D.M., Westbrook, R.F., 2017. A high-fat high-sugar diet-induced impairment in
Park, H., Kim, M., Shin, I.J., Park, J., Bang, S.Y., Yi, C., et al., 2018. place-recognition memory is reversible and training-dependent. Appetite 110, 61–71.
Woohwangcheongsimwon Prevents High-Fat Diet-Induced Memory Deficits and https://doi.org/10.1016/j.appet.2016.12.010.
Induces SIRT1 in Mice. J. Med. Food 21 (2), 167–173. https://doi.org/10.1089/jmf. Tran, D., Westbrook, R.F., 2018. Dietary effects on object recognition: the impact of high-
2017.3925. fat high-sugar diets on recollection and familiarity-based memory. J. Exp. Psychol.
Pathan, A.R., Gaikwad, A.B., Viswanad, B., Ramarao, P., 2008. Rosiglitazone attenuates Anim. Learn. Cogn. 44 (3), 217. https://doi.org/10.1037/xan0000170.
the cognitive deficits induced by high fat diet feeding in rats. Eur. J. Pharmacol. 589 Valladolid-Acebes, I., Stucchi, P., Cano, V., Fernandez-Alfonso, M.S., Merino, B., Gil-
(1-3), 176–179. https://doi.org/10.1016/j.ejphar.2008.06.016. Ortega, M., et al., 2011. High-fat diets impair spatial learning in the radial-armmaze
Piber, D., Nowacki, J., Mueller, S.C., Wingenfeld, K., Otte, C., 2018. Sex effects on spatial in mice. Neurobiol. Learn. Mem. 95 (1), 80–85. https://doi.org/10.1016/j.nlm.2010.
learning but not on spatial memory retrieval in healthy young adults. Behav. Brain 11.007.
Res. 336, 44–50. https://doi.org/10.1016/j.bbr.2017.08.034. Valladolid-Acebes, I., Fole, A., Martín, M., Morales, L., Cano, M.V., Ruiz-Gayo, M., Del
Popkin, B.M., Adair, L.S., Ng, S.W., 2012. Global nutrition transition and the pandemic of Olmo, N., 2013. Spatial memory impairment and changes in hippocampal mor-
obesity in developing countries. Nutr. Rev. 70 (1), 3–21. https://doi.org/10.1111/j. phology are triggered by high-fat diets in adolescent mice. Is there a role of leptin?
1753-4887.2011.00456.x. Neurobiol. Learning Memory 106, 18–25. https://doi.org/10.1016/j.nlm.2013.06.
Pratchayasakul, W., Sa-Nguanmoo, P., Sivasinprasasn, S., Pintana, H., Tawinvisan, R., 012.
Sripetchwandee, J., et al., 2015. Obesity accelerates cognitive decline by aggravating Vasileska, A., Rechkoska, G., 2012. Global and regional food consumption patterns and
mitochondrial dysfunction, insulin resistance and synaptic dysfunction under es- trends. Procedia-Social Behav. Sci. 44, 363–369. https://doi.org/10.1016/j.sbspro.
trogen-deprived conditions. Horm. Behav. 72, 68–77. https://doi.org/10.1016/j. 2012.05.040.
yhbeh.2015.04.023. Vorhees, C.V., Williams, M.T., 2006. Morris water maze: procedures for assessing spatial
Pugazhenthi, S., Qin, L., Reddy, P.H., 2017. Common neurodegenerative pathways in and related forms of learning and memory. Nat. Protoc. 1 (2), 848. https://doi.org/
obesity, diabetes, and Alzheimer’s disease. Biochimica et Biophysica Acta (BBA)- 10.1038/nprot.2006.116.

420
K.N. Abbott, et al. Neuroscience and Biobehavioral Reviews 107 (2019) 399–421

Vorhees, C.V., Williams, M.T., 2014. Assessing spatial learning and memory in rodents. Woodie, L., Blythe, S., 2018. The differential effects of high-fat and high-fructose diets on
ILAR J. 55 (2), 310–332. https://doi.org/10.1093/ilar/ilu013. physiology and behavior in male rats. Nutr. Neurosci. 21 (5), 328–336. https://doi.
Voyer, D., Postma, A., Brake, B., Imperato-McGinley, J., 2007. Gender differences in org/10.1080/1028415X.2017.1287834.
object location memory: a meta-analysis. Psychon. Bull. Rev. 14 (1), 23–38. https:// Wu, A., Molteni, R., Ying, Z., Gomez-Pinilla, F., 2003. A saturated-fat diet aggravates the
doi.org/10.3758/BF03194024. outcome of traumatic brain injury on hippocampal plasticity and cognitive function
Voyer, D., Voyer, S.D., Saint-Aubin, J., 2017. Sex differences in visual-spatial working by reducing brain-derived neurotrophic factor. Neuroscience 119 (2), 365–375.
memory: a meta-analysis. Psychon. Bull. Rev. 24 (2), 307–334. https://doi.org/10. https://doi.org/10.1016/S0306-4522(03)00154-4.
3758/s13423-016-1085-7. Xu, T.J., Reichelt, A.C., 2018. Sucrose or sucrose and caffeine differentially impact
Wang, W., Le, A.A., Hou, B., Lauterborn, J.C., Cox, C.D., Levin, E.R., et al., 2018. Memory- memory and anxiety-like behaviours, and alter hippocampal parvalbumin and dou-
related synaptic plasticity is sexually dimorphic in rodent hippocampus. J. Neurosci. blecortin. Neuropharmacology 137, 24–32. https://doi.org/10.1016/j.neuropharm.
38 (37), 7935–7951. https://doi.org/10.1523/JNEUROSCI.0801-18.2018. 2018.04.012.
Wechsler, D., Stone, C., 1987. Wechsler Memory Scale-Revised. Psychological Xu, N., Meng, H., Liu, T.Y., Feng, Y.L., Qi, Y., Zhang, D.H., Wang, H.L., 2018. Sterol O-
Corporation, New York, NY. acyltransferase 1 deficiency improves defective insulin signaling in the brains of mice
Wei, L., Yao, M., Zhao, Z., Jiang, H., Ge, S., 2018. High-fat diet aggravates postoperative fed a high-fat diet. Biochem. Biophys. Res. Commun. 499 (2), 105–111. https://doi.
cognitive dysfunction in aged mice. BMC Anesthesiol. 18 (1), 20. https://doi.org/10. org/10.1016/j.bbrc.2018.02.122.
1186/s12871-018-0482-z. Yagi, S., Galea, L.A., 2018. Sex differences in hippocampal cognition and neurogenesis.
West, D.B., Boozer, C.N., Moody, D.L., Atkinson, R.L., 1992. Dietary obesity in nine inbred Neuropsychopharmacology 1. https://doi.org/10.1038/s41386-018-0208-4.
mouse strains. Am. J. Physiology-Regulatory Integr. Comparative Physiol. 262 (6), Yeomans, M.R., 2017. Adverse effects of consuming high fat-sugar diets on cognition:
R1025–R1032. https://doi.org/10.1152/ajpregu.1992.262.6.R1025. implications for understanding obesity. Proc. Nutr. Soc. 76 (4), 455–465. https://doi.
West, D.B., Waguespack, J., McCollister, S., 1995. Dietary obesity in the mouse: inter- org/10.1017/S0029665117000805.
action of strain with diet composition. Am. J. Physiology-Regulatory Integr. Zhao, Q., Stafstrom, C.E., Fu, D.D., Hu, Y., Holmes, G.L., 2004. Detrimental effects of the
Comparative Physiol. 268 (3), R658–R665. https://doi.org/10.1152/ajpregu.1995. ketogenic diet on cognitive function in rats. Pediatr. Res. 55 (3), 498–506. https://
268.3.R658. doi.org/10.1203/01.pdr.0000112032.47575.d1.

421

You might also like